Download as pdf or txt
Download as pdf or txt
You are on page 1of 390

Green Chemistry and Sustainable Technology

Bhalchandra M. Bhanage
Masahiko Arai Editors

Transformation
and Utilization
of Carbon
Dioxide
Green Chemistry and Sustainable Technology

Series Editors

Prof. Liang-Nian He
State Key Laboratory of Elemento-Organic Chemistry, Nankai University, Tianjin,
China
Prof. Robin D. Rogers
Center for Green Manufacturing, Department of Chemistry, The University of
Alabama, Tuscaloosa, USA
Prof. Dangsheng Su
Shenyang National Laboratory for Materials Science, Institute of Metal Research,
Chinese Academy of Sciences, Shenyang, China
and
Department of Inorganic Chemistry, Fritz Haber Institute of the Max Planck
Society, Berlin, Germany
Prof. Pietro Tundo
Department of Environmental Sciences, Informatics and Statistics, Ca’ Foscari
University of Venice, Venice, Italy
Prof. Z. Conrad Zhang
Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian, China

For further volumes:


http://www.springer.com/series/11661
Green Chemistry and Sustainable Technology
Aims and Scope
The series Green Chemistry and Sustainable Technology aims to present cutting-edge
research and important advances in green chemistry, green chemical engineering
and sustainable industrial technology. The scope of coverage includes (but is not
limited to):
– Environmentally benign chemical synthesis and processes (green catalysis,
green solvents and reagents, atom-economy synthetic methods etc.)
– Green chemicals and energy produced from renewable resources (biomass,
carbon dioxide etc.)
– Novel materials and technologies for energy production and storage (biofuels
and bioenergies, hydrogen, fuel cells, solar cells, lithium-ion batteries etc.)
– Green chemical engineering processes (process integration, materials diversity,
energy saving, waste minimization, efficient separation processes etc.)
– Green technologies for environmental sustainability (carbon dioxide capture,
waste and harmful chemicals treatment, pollution prevention, environmental
redemption etc.)
The series Green Chemistry and Sustainable Technology is intended to provide an
accessible reference resource for postgraduate students, academic researchers and
industrial professionals who are interested in green chemistry and technologies for
sustainable development.
Bhalchandra M. Bhanage • Masahiko Arai
Editors

Transformation and
Utilization of Carbon
Dioxide
Editors
Bhalchandra M. Bhanage Masahiko Arai
Department of Chemistry Division of Chemical Process Engineering
Institute of Chemical Technology Faculty of Engineering, Hokkaido University
Mumbai, India Sapporo, Japan

ISSN 2196-6982 ISSN 2196-6990 (electronic)


ISBN 978-3-642-44987-1 ISBN 978-3-642-44988-8 (eBook)
DOI 10.1007/978-3-642-44988-8
Springer Heidelberg New York Dordrecht London
Library of Congress Control Number: 2014931416

© Springer-Verlag Berlin Heidelberg 2014


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or
information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed. Exempted from this legal reservation are brief excerpts
in connection with reviews or scholarly analysis or material supplied specifically for the purpose of being
entered and executed on a computer system, for exclusive use by the purchaser of the work. Duplication
of this publication or parts thereof is permitted only under the provisions of the Copyright Law of the
Publisher’s location, in its current version, and permission for use must always be obtained from
Springer. Permissions for use may be obtained through RightsLink at the Copyright Clearance Center.
Violations are liable to prosecution under the respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of
publication, neither the authors nor the editors nor the publisher can accept any legal responsibility for
any errors or omissions that may be made. The publisher makes no warranty, express or implied, with
respect to the material contained herein.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


Preface

Carbon dioxide is one of the green house gases and the emission and accumulation
of a huge amount of CO2 causes significant and negative effects on the global
environment. It is crucial, therefore, to take effective measures to reduce CO2
emission and treat the CO2 emitted. Carbon dioxide is a safe and abundant resource
of C and O and the production of useful chemicals from CO2 conforms to the Green
Chemistry principles. Carbon dioxide may replace harmful reactants like CO and
phosgene currently used in chemical reaction processes. Carbon dioxide is a stable
and less reactive molecule and, hence, the CO2 activation has been investigated so
far by using various catalysts and processes. The practical use of CO2 as a reactant
could also have a contribution to the reduction of the undesired CO2 in the
environment.
This book is dedicated to a wide-angle review of chemical transformation and
utilization of CO2 in 14 chapters. It is composed of four parts, which are chemical
(organic and inorganic), electro- and photo-chemical, and biochemical transformation
of CO2 to a number of organic and inorganic, small- and large-scale products. The last
two chapters deal with the application of CO2 as a promoter as well as a reactant in
organic synthetic reactions.
Part I, consisting of seven chapters, deals with the use of CO2 in chemical
reactions. In Chap. 1, Prof. Kühn and coworkers describe valorization of CO2 to
organic products using organocatalysts. Chapters 2 and 3 give several examples of
direct and indirect utilization of CO2 to valuable products using heterogeneous
catalysts. In Chap. 4, Prof. Beller and coworker review hydrogenation and related
reduction of CO2 using molecular catalysts. Professor Arena and coworkers
describe hydrogenation of CO2 to bulk chemicals like methanol and dimethyl
ether in Chap. 5. Professor Ross describes fundamentals and perspectives of syngas
production using CO2 reforming in Chap. 6. In Chap. 7, Prof. Rieger and coworkers
explain utilization of CO2 as a C-1 block for the synthesis of polycarbonates. These
chapters demonstrate the usefulness and potential of CO2 as a reactant.
Part II consists of four chapters dedicated to photocatalytic, electrochemical, and
inorganic reactions of CO2. Professor Krebs and coworker explore the possibilities
of using flexible substrates as basis for CO2 fixation and photocatalytic reduction in

v
vi Preface

Chap. 8. Chapter 9, contributed by Prof. Yamashita and co-workers, describes


reductive conversion of CO2 by using various photocatalyst materials. In Chap. 10,
Prof. Senboku explains electrochemical fixation of CO2. Chapter 11 of Prof. Baciocchi
and co-workers describes accelerated carbonation processes for CO2 capture, storage,
and utilization.
Part III is directed to biological reactions of CO2, consisting of Chap. 12 of Prof.
Das and coworker. This chapter summarizes the role of various biological processes
available for CO2 mitigation and discusses the use of terrestrial plants, algae,
cyanobacteria, carbonic anhydrase enzyme, and various other bacteria in CO2
sequestration.
Part IV deals with the reactions in/under dense phase CO2 in two chapters. In
Chap. 13, Prof. He and co-workers describe homogeneous catalysis promoted by
CO2. Chapter 14 explains the potential of multiphase catalytic reactions in/under
dense phase CO2, in which CO2 is utilized as a reaction promoter.
We believe/hope that this book will be useful to readers in various fields who
take practical and scientific interests in the current and future Transformation and
Utilization of Carbon Dioxide. All these chapters have been written by overseas
experts in the related fields. We would like to express our sincere thanks to the
authors for their excellent contributions to this book and their names are listed on
separate pages. We are also thankful to Springer for their support in publishing
this book.

Sapporo, Japan Masahiko Arai


Mumbai, India Bhalchandra M. Bhanage
Contents

Part I Chemical Reactions

1 Valorization of Carbon Dioxide to Organic Products


with Organocatalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Michael H. Anthofer, Michael E. Wilhelm, Mirza Cokoja,
and Fritz E. Kühn
2 Direct Transformation of Carbon Dioxide to Value-Added
Products over Heterogeneous Catalysts . . . . . . . . . . . . . . . . . . . . . 39
Shin-ichiro Fujita, Masahiko Arai, and Bhalchandra M. Bhanage
3 Indirect Utilisation of Carbon Dioxide in Organic Synthesis
for Valuable Chemicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
Rahul A. Watile, Bhalchandra M. Bhanage, Shin-ichiro Fujita,
and Masahiko Arai
4 Hydrogenation and Related Reductions of Carbon Dioxide
with Molecular Catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Carolin Ziebart and Matthias Beller
5 Latest Advances in the Catalytic Hydrogenation
of Carbon Dioxide to Methanol/Dimethylether . . . . . . . . . . . . . . . . 103
Francesco Arena, Giovanni Mezzatesta, Lorenzo Spadaro,
and Giuseppe Trunfio
6 Syngas Production Using Carbon Dioxide Reforming:
Fundamentals and Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
Julian R.H. Ross
7 Carbon Dioxide as C-1 Block for the Synthesis
of Polycarbonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
Peter T. Altenbuchner, Stefan Kissling, and Bernhard Rieger

vii
viii Contents

Part II Photocatalytic, Electrochemical and Inorganic Reactions

8 Fixation of Carbon Dioxide Using Molecular Reactions


on Flexible Substrates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
Jacob Jensen and Frederik C. Krebs
9 Reductive Conversion of Carbon Dioxide Using Various
Photocatalyst Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
Kojirou Fuku, Kohsuke Mori, and Hiromi Yamashita
10 Electrochemical Fixation of Carbon Dioxide . . . . . . . . . . . . . . . . . . 245
Hisanori Senboku
11 Accelerated Carbonation Processes for Carbon Dioxide
Capture, Storage and Utilisation . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
Renato Baciocchi, Giulia Costa, and Daniela Zingaretti

Part III Biological Reactions

12 Carbon Dioxide Sequestration by Biological Processes . . . . . . . . . . 303


Kanhaiya Kumar and Debabrata Das

Part IV Reactions in/Under Dense Phase Carbon Dioxide

13 Homogeneous Catalysis Promoted by Carbon Dioxide . . . . . . . . . . 337


Ran Ma, Zhen-Feng Diao, Zhen-Zhen Yang, and Liang-Nian He
14 Multiphase Catalytic Reactions in/Under Dense-Phase
Carbon Dioxide: Utilization of Carbon Dioxide
as a Reaction Promoter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369
Hiroshi Yoshida, Shin-ichiro Fujita, Masahiko Arai,
and Bhalchandra M. Bhanage
Contributors

Peter T. Altenbuchner Technische Universität München, WACKER-Lehrstuhl


für Makromolekulare Chemie, Garching, Germany
Michael H. Anthofer Catalysis Research Center, Technische Universität
München, München, Germany
Masahiko Arai Division of Chemical Process Engineering, Faculty of Engineering,
Hokkaido University, Sapporo, Japan
Francesco Arena Department of Electronic Engineering, Industrial Chemistry
and Engineering, University of Messina, Messina, Italy
Renato Baciocchi Laboratory of Environmental Engineering, Department of Civil
Engineering and Computer Science Engineering, University of Rome “Tor
Vergata”, Rome, Italy
Matthias Beller Leibniz-Institut für Katalyse Eingetragener Verein an der
Universität Rostock, Rostock, Germany
Bhalchandra M. Bhanage Department of Chemistry, Institute of Chemical
Technology, Mumbai, India
Mirza Cokoja Catalysis Research Center, Technische Universität München,
München, Germany
Giulia Costa Laboratory of Environmental Engineering, Department of Civil
Engineering and Computer Science Engineering, University of Rome “Tor Vergata”,
Rome, Italy
Debabrata Das Department of Biotechnology, Indian Institute of Technology,
Kharagpur, India
Zhen-Feng Diao State Key Laboratory of Elemento-Organic Chemistry, Nankai
University, Tianjin, People’s Republic of China

ix
x Contributors

Shin-ichiro Fujita Division of Chemical Process Engineering, Faculty of


Engineering, Hokkaido University, Sapporo, Japan
Kojirou Fuku Division of Materials and Manufacturing Science, Graduate School
of Engineering, Osaka University, Suita, Osaka, Japan
Elements Strategy Initiative for Catalysts Batteries ESICB, Kyoto University,
Katsura, Kyoto, Japan
Liang-Nian He State Key Laboratory of Elemento-Organic Chemistry, Nankai
University, Tianjin, People’s Republic of China
Jacob Jensen Department of Energy Conversion and Storage, Technical University
of Denmark, Roskilde, Denmark
Stefan Kissling Technische Universität München, WACKER-Lehrstuhl für
Makromolekulare Chemie, Garching, Germany
Frederik C. Krebs Department of Energy Conversion and Storage, Technical
University of Denmark, Roskilde, Denmark
Fritz E. Kühn Catalysis Research Center, Technische Universität München,
München, Germany
Kanhaiya Kumar Department of Biotechnology, Indian Institute of Technology,
Kharagpur, India
Ran Ma State Key Laboratory of Elemento-Organic Chemistry, Nankai University,
Tianjin, People’s Republic of China
Giovanni Mezzatesta Department of Electronic Engineering, Industrial Chemistry
and Engineering, University of Messina, Messina, Italy
Kohsuke Mori Division of Materials and Manufacturing Science, Graduate
School of Engineering, Osaka University, Suita, Osaka, Japan
Elements Strategy Initiative for Catalysts Batteries ESICB, Kyoto University,
Katsura, Kyoto, Japan
Bernhard Rieger Technische Universität München, WACKER-Lehrstuhl für
Makromolekulare Chemie, Garching, Germany
Julian R.H. Ross University of Limerick, Limerick, Ireland
Hisanori Senboku Division of Chemical Process Engineering, Faculty of
Engineering, Hokkaido University, Sapporo, Hokkaido, Japan
Lorenzo Spadaro Department of Electronic Engineering, Industrial Chemistry
and Engineering, University of Messina, Messina, Italy
Giuseppe Trunfio Department of Electronic Engineering, Industrial Chemistry
and Engineering, University of Messina, Messina, Italy
Contributors xi

Rahul A. Watile Department of Chemistry, Institute of Chemical Technology,


Mumbai, India
Michael E. Wilhelm Catalysis Research Center, Technische Universität
München, München, Germany
Hiromi Yamashita Division of Materials and Manufacturing Science, Graduate
School of Engineering, Osaka University, Suita, Osaka, Japan
Elements Strategy Initiative for Catalysts Batteries ESICB, Kyoto University,
Katsura, Kyoto, Japan
Zhen-Zhen Yang State Key Laboratory of Elemento-Organic Chemistry, Nankai
University, Tianjin, People’s Republic of China
Hiroshi Yoshida Division of Chemical Process Engineering, Faculty of Engineering,
Hokkaido University, Sapporo, Japan
Carolin Ziebart Leibniz-Institut für Katalyse Eingetragener Verein an der
Universität Rostock, Rostock, Germany
Daniela Zingaretti Laboratory of Environmental Engineering, Department
of Civil Engineering and Computer Science Engineering, University of Rome
“Tor Vergata”, Rome, Italy
Part I
Chemical Reactions
Chapter 1
Valorization of Carbon Dioxide to Organic
Products with Organocatalysts

Michael H. Anthofer, Michael E. Wilhelm, Mirza Cokoja,


and Fritz E. Kühn

1.1 Introduction

Humankind is facing an increasing problem with energy supply. Society must meet
the needs of the steadily increasing world population [1], and additionally, indus-
trially emerging countries exhibit a rising energy demand in order to improve their
living standard. Today, the main energy sources are fossil carriers such as petro-
leum and, to a minor extent, coal [2]. Other sources are either ecologically or
politically problematic, such as nuclear power, or they are not yet efficient enough,
depend on climatic fluctuations, or are difficult to store (solar and wind power).
Hence, if the world continues to rely on carbon compounds as energy carriers, with
the energy stored in chemical bonds, other renewable resources have to be consid-
ered. So far, however, the main alternative carbon feedstock – natural gas, carbon
dioxide, and biomass (starch, cellulose, lignin) – have primarily been used for direct
combustion, and their transformation into value-added chemicals is considered as
notoriously difficult due to their chemical inertness [3–6]. For this reason, catalysis
research is increasingly devoted to finding ways to develop high-performance
catalysts for the valorization of renewable carbon feedstocks [7–9]. Numerous
approaches have already been presented in the literature. However, they are often
associated with a significant carbon footprint, which is thus rendering these
approaches useless, at least on a larger scale [10–14].
Carbon dioxide is in a special spotlight as a renewable C1-feedstock, since it is
ubiquitous and readily available, either from decomposition of organic matter or
artificially. Whereas CO2 can be transformed to carbohydrates in nature – which is
the basis of life on Earth – from a chemical (industrial) point of view, CO2 appeared

Michael H. Anthofer and Michael. E. Wilhelm equally contributed to this article.


M.H. Anthofer • M.E. Wilhelm • M. Cokoja (*) • F.E. Kühn (*)
Catalysis Research Center, Technische Universität München, Ernst-Otto-Fischer-Straße 1,
D-85747 Garching bei München, Germany
e-mail: mirza.cokoja@ch.tum.de; fritz.kuehn@ch.tum.de

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 3
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_1,
© Springer-Verlag Berlin Heidelberg 2014
4 M.H. Anthofer et al.

to be a dead end for all organic compounds. The idea behind the use of CO2 as
feedstock for chemicals and fuels is nevertheless attractive: the direct energy supply
via combustion of carbonaceous energy carriers produces CO2, which could be
converted back into the energy carrier, thereby leading to an artificial carbon cycle,
leading to a recycling of CO2 and balancing the anthropogenic emission into the
atmosphere [15]. Evidently, the activation of CO2 can only be realized using
catalysts (and – in many cases additional – cheap energy sources). In this case,
the classic heterogeneous catalysts fail to convert CO2 at mild temperatures, which
would warrant a multiton plant scale. On the other hand, homogeneous catalysts
usually operate at significantly milder conditions [16–18]. From all possible trans-
formations of CO2 to C1 or higher chemicals with molecular catalysts in solution,
the most promising reactions in terms of energy balance, catalyst performance,
reusability, and scale-up perspective would be:
• Reduction to CO
• Hydrogenation with H2 to formate, methanol, or methane
• Formation of (poly)carbonate
The reduction reactions appear particularly attractive, since the products are all
significantly more reactive than CO2 or can themselves be used as energy carriers.
However, these processes are to date either not efficient enough, or they involve
hydrogen, which is currently expensive (steam reforming process) and therefore an
undesired reactant. As long as there are no promising approaches to access H2 from
water cheaply, e.g., photochemically, these reactions are prohibitive for industry.
Contrasting this, the synthesis of carbonates and polycarbonates from CO2 and
epoxides is a well-known and established procedure using various high-performance
metal catalysts [19–25]. Yet, for all reactions, there is plenty of room for improve-
ment of the catalyst design, efficiency, and reaction conditions in order to establish a
process for the valorization of CO2 (in the multimillion ton scale).
In comparison to molecular metal catalysts, metal-free catalysts are not so
widespread. However, the application of organocatalysts would – in most cases –
easily outcompete metal catalysts in many regards, not only in price or stability but
also in terms of “green chemistry,” avoiding heavy metal waste or intricate catalyst
recovery. This chapter presents a comprehensive overview of catalytic transforma-
tions of CO2 to organic products – most notably carbonates – employing molecular
metal-free catalysts reported in the literature to date.

1.2 Ionic Compounds as Catalysts for the Green


Synthesis of Cyclic Carbonates from Epoxides
and Carbon Dioxide

The reaction with epoxides to cyclic carbonates (Scheme 1.1) is one of the best-
studied methods for the chemical fixation of CO2 in homogeneous phase. The most
important cyclic carbonates are ethylene carbonate (EC), propylene carbonate (PC),
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 5

Scheme 1.1 Conversion of O


carbon dioxide and
epoxides into cyclic O O O
+ CO2
carbonates R1 R2
R1 R2

O O
HO R X
Br Bu4N

O O

O O NBu4 O O O H O H O R
Br Bu4N
Br X R O
X

H O
CO2 O NBu4 O
Br CO2 X R

Scheme 1.2 Proposed mechanisms for the chemical fixation of CO2 with EO using TBAB as
catalyst and a proposed cooperative mechanism for hydroxyl-functionalized ionic liquids (HFILs)

and styrene carbonate (SC). Five-membered cyclic carbonates can be used as polar
aprotic solvents, electrolytes in secondary batteries, valuable precursors for poly-
carbonate and polyurethane materials, and intermediates in organic synthesis
[26]. The first successful approach was carried out by Calo et al. by dissolving
propylene oxide (PO) in tetrabutylammonium bromide (TBAB) or tetrabuty-
lammonium iodide (TBAI) at atmospheric CO2 pressure to yield PC [27]. The
observed activity of the catalytic system decreases in order of TBAI > TBAB due
to the higher nucleophilicity of iodide in comparison to bromide. In addition, an
increase of chain length from tetraethylammonium bromide (TEAB) to TBAB
results in a higher yield of PC. After the reaction is completed, the pure cyclic
carbonate can be easily separated by vacuum distillation or extraction with ethyl
acetate, which allows the recycling of the ammonium salt. For getting a better
understanding of the reaction mechanism, investigations toward the cycloaddition
of epoxides with CO2 using TBAB as catalyst were studied in various reports.
Proposed mechanisms are illustrated in Scheme 1.2 [27, 28].
The bromide anion performs a nucleophilic attack to form an oxyanion species.
This species reacts with CO2 to result in cyclic carbonate formation after intramo-
lecular cyclic elimination. Further experimental and theoretical studies of the
catalytic behavior, in dependency of the used anion with a consistent cation,
show a higher activity in the order Cl > Br > I, which was not in accord with
the calculated relative energy of the rate-determining step [28]. Nevertheless,
lowering of the activation energies and thus higher activity could be confirmed
experimentally and theoretically.
6 M.H. Anthofer et al.

Tetraalkylammonium salts Imidazolium salts Guanidinium salts


R

N X X
N X
R N N N N
R R1 R2
R
H H

TEAB: R=C2H5; X=Br Cations:


HTBDCl: X=Cl
TBAB: R=C4H9; X=Br OMIm: R1=C8H17; R2=CH3 HTBDBr: X=Br
TBAI: R=C4H9; X=I BMIm: R1=C4H9; R2=CH3
MMIm: R1=CH3 R2=CH3
Anions:
BF4, PF6, Cl, Br, I

Fig. 1.1 Representative tetraalkylammonium-, imidazolium-, and guanidinium-based ionic liquids

The first quantitative conversion of propylene oxide (PO) to propylene carbonate


(PC) was observed in 2001 by Peng and Deng, using 2.5 mol-% of 1-butyl-3-
methylimidazolium tetrafluoroborate (BMImBF4) under 25 bar CO2 pressure at
110  C after 6 h (TOF ¼ 6.6 h1) [29]. The group investigated the influence of
cation and anion of the ionic liquid (IL) and asserted an increase in the catalytic
activity in order imidazolium > pyridyl for the cation and PF6 > Cl > BF4 for the
anion. In later works, the catalytic activity of several ammonium halides and
imidazolium-based ILs could be further improved by microwave irradiation instead
of heating [30]. A representative overview of different ILs active as organocatalysts
for the fixation of CO2 is given in Fig. 1.1.
In the next step, imidazolium-based ILs were used as catalyst with supercritical
CO2 (scCO2) as solvent and reactant. Therefore, 1-octyl-3-methylimidazolium
tetrafluoroborate (OMImBF4) was found to be an excellent catalyst under super-
critical conditions. Nearly 100 % yield and selectivity of PC was determined at
140 bar CO2 pressure and 100  C after 5 min (TOF ¼ 517 h1) [31]. The TOF
value is nearly 80 times faster than previously reported [29]. Ikushima
et al. detected an increase of the yield of PC by increasing the chain length of the
alkyl side chain of the imidazolium cation from C2 to C8, due to higher solubility of
CO2 and PO in the IL phase [31]. Furthermore, theoretical studies indicate coop-
erative interactions between cation and anion, which stabilize intermediates and
transition states through hydrogen bonding. Therefore, the ring-opening of PO as
the rate-determining step is facilitated, leading to enhanced catalytic activity
[32]. By comparison of 1-butyl-3-methylimidazolium chloride (BMImCl) and
1-butyl-3-methylimidazolium bromide (BMImBr) for the synthesis of PC, TOFs
of 5 h1 (BMImCl) and 37 h1 (BMImBr) were found at 100  C, 35 bar CO2
pressure after 1 h [33]. Hence, the activity of the catalyst decreases in order of
Br > Cl.
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 7

Fig. 1.2 BMPPPA∙3Br as


reaction medium and
efficient organocatalyst for
N N N
CO2 fixation into cyclic
carbonates
3•Br
N N

BMPPPA•3Br

Fig. 1.3 Betaine-based


salts and TMAP as X N COOH NMe3
bifunctional
organocatalysts bearing a
betaine framework HBetCl: X=Cl
HBetBr: X=Br O
HBetI: X=I
HBetBF4: X=BF4
HBetPF6: X=PF6 TMAP

Wong et al. prepared the new IL BMPPPA · 3Br (Fig. 1.2) with a Lewis acidic
cation, which is a promising catalyst for the activation of epoxides in the cycload-
dition with CO2. This IL catalyzes the formation of PC with a yield and selectivity
of 99 % at 80  C, 40 bar CO2 pressure after 3 h and could be recycled for more
than ten times without any loss of activity. Small amounts of water (2 % w/w) are
needed to liquefy the IL at 80  C in order to function also as reaction medium [34].
Task-specific ILs (TSILs) bearing a functional group, allowing the activation of PO
or CO2 or both, are the next generation of ILs for the formation of cyclic carbonates.
First, a series of hydroxyl-functionalized ionic liquids (HFILs) was developed, char-
acterized, and found to be efficient and reusable catalysts for the synthesis of cyclic
carbonates with high selectivity in the absence of any cocatalyst and any cosolvents
[35]. Hydroxyl groups in the catalyst were proposed to polarize the C–O bond via
hydrogen bonding and therefore facilitate the nucleophilic attack of the halide anion,
as illustrated in Scheme 1.2. 1-(2-hydroxyethyl)-3-methylimidazolium bromide
(HEMImBr) was most efficient for the synthesis of PC with 99 % conversion of PO
and a selectivity of 99.8 % under the optimized reaction conditions (catalyst loading:
1.6 mol-%, 125  C, 20 bar CO2, 1 h).
Similar to hydroxyl groups, carboxylic acid groups as hydrogen bond donors can
also accelerate the ring-opening reaction of the epoxide. Through direct protonation
of anhydrous betaine, a series of bifunctional catalysts was synthesized, allowing
the polarization of the epoxide through hydrogen bonding and at the same time the
ring-opening of the epoxide with the respective anion (Fig. 1.3) [36].
The catalytic activity for the formation of PC increases in order HBetI > HBetCl >
HBetBr > HBetBF4 > HBetPF6. Under catalytic conditions, 2.5 mol-% HBetI
8 M.H. Anthofer et al.

(80 bar CO2 pressure, 140  C, 8 h) result in a yield of 98 % PC and the catalyst
was recycled for at least three times. Based on these results, organocatalysts
bearing an ammonium betaine moiety have been developed as a new catalytic
motif for the activation of carbon dioxide and epoxides to produce cyclic carbon-
ates under milder reaction conditions [37]. After screening several catalysts,
3-trimethylammoniumphenolate (TMAP) was shown to be the most active for the
conversion of PO to PC with a yield of 99 % under relatively mild reaction
conditions (catalyst loading: 3 mol-%, 10 bar CO2 pressure, 120  C, 24 h). Addi-
tionally, TMAP is capable to form a CO2 adduct at 15  C which is also promising for
carbon capture and storage (CCS). Another attempt for a bifunctional catalyst
system was using α-amino acids in scCO2 due to the same background as mentioned
before for betaine. After a screening, L-histidine turned out to be the most active
catalyst leading to the corresponding cyclic carbonates in good to excellent
yields with a catalyst loading of 0.8 mol-% at 130  C under 80 bar CO2 [38]. Natural
alpha-amino acids-derived ionic liquids, comprising BMIm cation and amino acid
anion, were found to be efficient catalysts for the coupling of various epoxides and
CO2. Cyclic carbonates could be produced in good yields and selectivity without any
additional organic solvent or halogen [39]. Using a modified amino acid IL with
triethylamine (NEt3) as cocatalyst, the catalytic activity of the L-proline-based ionic
liquid was greatly enhanced. Up to 97 % isolated yield of cyclic carbonate was
achieved at 90  C under atmospheric CO2 pressure without organic solvent and
metal component [40].
Another way for the fixation of CO2 in cyclic carbonates are CO2 adducts
of N-heterocyclic carbenes (NHCs). Therefore, the thermal stable 1,3-bis
(2,6-diisopropylphenyl)imidazolium-2-carboxylate (IPr-CO2) was found to cata-
lyze the reaction of PO with CO2 to PC. Under the optimized reaction conditions
(catalyst loading: 0.5 mol-%, 2 mL CH2Cl2, 20 bar CO2, 120  C, 24 h), a
quantitative conversion of PO with a selectivity 99 % for PC was observed
[41]. Ikariya et al. also developed a series of different NHC-CO2 adducts which
were active toward the cycloaddition of CO2 with epoxides. With a catalyst loading
of 5.0 mol-%, styrene oxide was converted with a selectivity 99 % to the
corresponding cyclic carbonate in a yield of 98 % under optimized reaction
conditions (45 bar CO2 pressure, 100  C, 24 h) [42].
Cheng et al. found a single Lewis base catalyst for the formation of cyclic
carbonates from CO2 and terminal epoxides in water without suffering selectivity
[43]. It is proposed that water and the anion of the Lewis base play a synergistic
effect in accelerating the ring-opening reaction. Among the investigated catalysts,
triphenylbutylphosphonium iodide (PPh3BuI) was the most active with a yield of
95 % and a selectivity of 98 % for the conversion of ethylene oxide (EO) to ethylene
carbonate (EC) under optimized reaction conditions (0.5 mol-% PPh3BuI,
33.5 mol-% water, 110  C, 25 bar CO2 pressure, 0.5 h). The catalyst could be
recycled for at least ten times without significant loss of activity.
A series of easily prepared Lewis basic ILs were developed for cyclic carbonate
synthesis from epoxide and CO2 without utilization of any organic solvents or
additives (Fig. 1.4) [44]. The presence of a tertiary nitrogen atom in the cation has
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 9

H H
N N N
n
N X
N N N X
X
Cations: N
C4DABCO: n=3
C8DABCO: n=7 HDBUCl: X=Cl HHMTACl: X=Cl
Anions: HDBUOAc: X=OAc
X=Cl, Br, OH, BF4, PF6, NTf2

Fig. 1.4 Lewis basic ionic liquids

Fig. 1.5 Imidazolium- N N


HO n OH N N OH
based acid–base n n
bifunctional catalysts O O O
X X
(ABBCs)
DCC1ImCl: n=1; X=Cl CC1ImCl: n=1; X=Cl
DCC1ImBr: n=1; X=Br CC1ImBr: n=1; X=Br
DCC4ImCl: n=3; X=Cl CC4ImBr: n=3; X=Br
DCC4ImBr: n=3; X=Br

the potential to form a carbamate species as an activated form of CO2. Additionally,


the epoxide could be activated through hydrogen bonding of the proton attached to
the ammonium nitrogen atom of the IL. In dependence of the cation, the catalytic
activity decreases in order of HDBU+ > HTBD+ ~ OMIm+ > C4DABCO+ ~
C8DABCO+ > BMIm+ > HHMTA+. After 2 h, 97 % yield and a selectivity
99 % toward the formation of PC were attained, when 1.0 mol-%
1,8-diazabicyclo [5.4.0]undec-7-enium chloride ([HDBU]Cl) was used as a catalyst
at 140  C under 10 bar CO2 pressure. Furthermore, the catalyst could be recycled
over five times without appreciable loss of catalytic activity. Also, hydroiodides of
amidines catalyze the reaction of CO2 and epoxides. Under ambient pressure and
temperature, the corresponding cyclic carbonates were obtained in moderate to high
yields [45]. Furthermore, hydroiodides of secondary and primary amines were found
to be efficient catalysts for PC synthesis [46]. Detailed investigation showed that the
catalytic activity is strongly dependent on the counter anions of the ammonium salts.
Iodide catalyzes the reaction efficiently, whereas bromide and chloride show almost
no activity. In addition, the activity of the catalyst increases with increasing bulk-
iness of the substituents on the ammonium nitrogen atoms, since PC synthesis
becomes more efficient as the parent amines become more basic. After performing
a catalyst screening, dicyclohexylammonium iodide (DCAI) was determined to
be the most active catalyst. SC was successfully synthesized out of SO by using
5 mol-% DCAI dissolved in 1-methyl-2-pyrrolidone (NMP) at 45  C under
atmospheric CO2 pressure after 24 h in a yield of 85 % and a selectivity 99 %.
New acid–base bifunctional catalysts (ABBCs), containing one or two Brønsted
acidic sites in the imidazolium-based cationic part and a Lewis basic site in the anionic
part, were determined as efficient catalysts for the synthesis of cyclic carbonates
without the use of additional cocatalyst or cosolvent (Fig. 1.5) [47]. Imidazolium was
10 M.H. Anthofer et al.

Scheme 1.3 Preparation of H


amidinium halides R N NMe2 R N NMe2 X
HX

C6-A: R=C6H13 C6-ACl: R=C6H13; X=Cl


C6-ABr: R=C6H13; X=Br
C6-AI: R=C6H13; X=I

used as cationic structural motif due to a decrease of catalytic activity in order of


imidazolium- > bipyridinium- > isonicotinic-based ABBCs. The catalytic cycle is
proposed to be the same as it is for HFILs and illustrated in Scheme 1.2.
Among all catalysts, 1,3-(bis-3-carboxypropyl)imidazolium bromide DCC4ImBr
was the most active and could be recycled for five times without loss of activity.
Almost quantitative yield of PC can be achieved in 1 h at 100  C under 20 bar CO2.
Park et al. synthesized ILs with carboxylic acid moieties in their backbones
and observed a higher activity for the cycloaddition reaction of CO2 with various
epoxides for ABBCs, compared to HFILs and ILs without any functional
groups [48].
Tassaing and coworkers investigated the catalytic cycloaddition of PO to CO2 at
80  C and 80 bar CO2 pressure via in situ FTIR and Raman spectroscopy using the
primarily tetraalkylammonium, imidazolium, and guanidinium salts mentioned in
Fig. 1.1 [49]. Among all investigated ILs, 1,5,7-triazabicyclo[4.4.0]dec-5-enium
bromide (HTBDBr) and 1-methyl-3-methylimidazolium iodide (MMImI) showed
the highest activity. The spectroscopic results show the possibility of activating the
epoxide through hydrogen bonding and stabilization of the oxyanion intermediate.
This is comparable to other works and fits to the illustrated mechanism in
Scheme 1.2 [35, 44, 48, 50–52]. Therefore, SO was converted to SC with a yield
of 87 % using 1 mol-% HTBDBr at 80  C under 10 bar CO2 pressure after 20 h. If
the reaction is carried out at 80 bar CO2, the yield of SC is only about 88 %. By
pressurizing the reaction with less than 10 bar CO2, the yield decreases to 65 %.
Nevertheless, no recycling experiments were carried out to show the efficiency of
the catalytic system.
The IL-catalyzed reaction of PO and CO2 to PC was also studied in a microreactor
using 2-hydroxyethyltributylammonium bromide (HETBABr) as catalyst. After a
parameter variation, the results show that the residence time can be dramatically
reduced from several hours in a conventional stirred reactor to about 10 s in a
microreactor. Almost quantitative yield of PC can be reached at 35 bar at a residence
time of 14 s. The TOF value varies in the range of 3,000–14,000 h1 compared to
60 h1 in the conventional stirred reactor. This work opens the possibility for the
commercialization of the metal-free conversion of CO2 to cyclic carbonates [53].
Amidinium halide catalysts were synthesized (see Scheme 1.3) and employed in
the synthesis of cyclic carbonates using an in situ-prepared, reversible, room-
temperature ionic liquid (RTIL), amidinium carbamates, as solvent [54].
Amidinium carbamates (C6-A-CO2) were prepared by exposing equimolar
mixtures of C6-A and a primary amine to CO2 gas. 2 mol-% C6-AI in C6-A-CO2
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 11

Cl
O C O O O
H R O H R O H R
R , CO2 R N O
N P NR Cl N P N R P NR
N R N N
N N N

PA-methyl: R=CH3
PA-p-methoxybenzyl: R=CH2-C6H4-OCH3
PA-neo-pentyl: R=CH2-C(CH3)3

O
O O Cl Cl
O
O O O
H R O H R
R O
N P NR P
NR
N R N N
N N

Scheme 1.4 Proposed mechanism for the azaphosphatrane derivatives catalyzed cycloaddition of
EO and CO2

at 50  C under 25 bar CO2 pressure lead to a yield of 99 % of SC after 8 h. Product


and any unreacted epoxide were extracted easily upon addition of an immiscible,
lower density liquid to the RTIL reaction mixture. This process was repeated three
times with the same RTIL without any obvious decrease in catalytic activity. This
system is an attractive alternative for preparing cyclic carbonates and CO2 storage
due to the reusability and the rather low temperatures and CO2 pressures.
Latest research efforts lead to azaphosphatrane (AP) derivatives as a new class
of catalysts [55]. AP can be a tunable alternative to known ammonium halides, but
their catalytic behavior strongly depends on their substitution pattern cause of the
proposed mechanism. The authors propose an activation of the epoxide through
hydrogen bonding of the phosphonium cation and an insertion of CO2 into the P–N
bond. The formed tricyclic phosphoryl-carbamate species is very sensitive. Bulkier
substituents prevent degradation of the catalyst and lead to a higher activity
(Scheme 1.4). AP-neo-pentyl with the most bulky substituents is the most stable
and active catalyst in this work. A kinetic experiment with 1.0 mol-% AP-neo-
pentyl leads to 50 % yield of SC in toluene as solvent at 100  C under atmospheric
CO2 pressure after 7 h.

1.3 Formation of Cyclic Carbonates Catalyzed


by Immobilized Ionic Liquids

The separation of homogeneous catalysts from the reaction mixture is often asso-
ciated with complicated procedures, which possibly prevent the industrial applica-
tion. In order to facilitate the recycling of ionic liquids as metal-free catalysts for
12 M.H. Anthofer et al.

OR
+ RO R⬘ IL+ X- method 1
Si(OEt)4 Si
RO R⬘ IL+ X-

OR
RO
OH + Si R⬘ IL+ X- method 2
RO R⬘ IL+ X-

R⬘ X + IL R⬘ IL+ X- method 3

IL = NR3, PR3, RIm, Rpy, guanidnium X = Cl, Br, I

Scheme 1.5 Methods for the covalent immobilization of ILs on supporting material

the cycloaddition of CO2, the immobilization on solid carriers leading to supported


ionic liquid phase (SILP) materials is a promising concept. The obtained heteroge-
neous catalysts do not only offer advantages with regard to catalyst separation and
reusability but also the selectivity and reaction rate can be increased through the
proper use of the supporting material.
Impregnation of ILs via physical adsorption on carrier materials illustrates one
route for the preparation of SILPs. The synthesis of this type of catalysts is carried
out by dispersing the support materials in the respective solvent, adding the desired
amount of IL followed by evaporation of the solvent. By using this method, Wang
et al. prepared silica gel-supported tetraalkylammonium and imidazolium salts,
which gave full conversion of PO with a selectivity up to 98 % for PC under
supercritical conditions (150–160  C, 80 bar CO2) [56, 57]. Zhu et al. immobilized
choline chloride/urea on molecular sieves and achieved quantitative conversion of
PO after 5 h at 110  C [58]. In all of these cases, the catalytic activity of the
respective ILs is nearly unaffected by their immobilization. However, the resulting
catalysts can be separated from the reaction mixture by simple filtration and are
recyclable for several times.
The covalent immobilization of ILs on amorphous or mesoporous silica and
zeolites represents another possibility for the synthesis of SILP catalysts. This can
be accomplished by the cocondensation of trialkoxysilylalkyl-functionalized ILs and
triethoxysilane (Method 1, Scheme 1.5) [59, 60] or the reaction of these ILs with
hydroxyl groups on the surface of the supporting material (Method 2, Scheme 1.5)
[61, 62]. In addition to that, the quaternization of terminally monohalogenated alkyl-
functionalized carrier materials can also be used to produce SILP materials (Method
3, Scheme 1.5) [63, 64]. The reaction conditions for the formation of cyclic carbonates
with the corresponding catalysts and supporting materials together with the synthetic
methods are given in Table 1.1.
These covalently immobilized IL catalysts are not only easily recyclable from
the reaction mixture, but also show a higher catalytic activity than the homogeneous
analogs, which is not convenient to other catalytic reactions. This derives from a
synergistic effect from the silanol groups on the surface of the silica material, which
Table 1.1 Formation of cyclic carbonates with silica-based SILP catalysts
IL Support material Method Substrate Cat. loading T [ C] p (CO2) [bar] t [h] Conversion [%] Selectivity [%] Ref.
a
MImCl Silica gel 1 PO 17.6 wt.-% 120 8.0 86 Not given [60]
BImBr Silica gel 1 AGE 11.0 wt.-% 110 35.5 6.0 99 93 [65]
BImI Silica gel 2 AGE 11.0 wt.-% 110 7.6 3.0 78 100 [68]
MImOH Silica gel 2 PO 1.8 mol-% 120 20.0 4.0 95 100 [61]
EtImBr MCM-41 3 AGE 11.0 wt.-% 110 21.0 6.0 99 95 [66]
CC3ImBr Silica gel 2 PO 21.5 wt.-% 110 16.0 3.0 100 99 [69]
Br-EImBr MCM-41 3 SO 8.3 wt.-% 140 40.0 4.0 100 99 [70]
EImBr SBA-15 1 PGE 8.9 wt.-% 140b 9.7 0.25 91 100 [67]
PBGCl Silica gel 1 PO 23.1 wt.-% 120 45.0 4.0 100 >99 [59]
4-pyr-pyI Silica gel 3 SO 0.8 mol-% 100 10.0 20.5 91 98 [71]
PBu3I Silica gel 3 PO 1.0 mol-% 100 100.0 1 100 100 [63]
PPh3Br Silica gel 2 PO 1.0 mol-% 90 10.0 6 100 99 [62]
NEt3Cl MCM-41 3 AGE 22.4 wt.-% 110 17.6 6 73 99 [72]
HETABr SBA-15 2 PO 1.0 mol-% 110 20.0 2 99 99 [73]
Br-E 3-(2-Bromoethyl), PBG pentabutyl-guaniddinium, 4-pyr-py 4-pyrrolidinopyridinium, HOEtTA 4-(2-Hydroxy-ethyl)-1,2,4-triazolium, AGE allyl
glycidyl ether
a
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts

PO/CO2 ¼ 2/1
b
100 W
13
14 M.H. Anthofer et al.

n R = nBu; X = Cl; a
n R = Et, C2H4OH, CH2COOH; X = Br; b
N DVB n DVB: 1,4-divinylbenzene
N or N
+ - AIBN X- AIBN: azobisisobutyronitrile
N X N+ X- n
R N+
R R

R= nBu, Ph; c
N EGDMA VCC2MImCl; d
or CLPN
N+ Cl- AIBN EGDMA: ethylene glycol dimethacrylate
CLPN: cross-linked polymeric nanoparticle
PR3 Cl- MeO
+
O

Scheme 1.6 Preparation of SILP catalysts by radical polymerization

activate the epoxide, thus facilitating the nucleophilic attack [63]. The influence of
the catalyst loading on the reactivity of the system was examined and it was pointed
out that higher amount of immobilized IL and more bulky catalysts lead to a
decrease in the reactivity. This is caused by the reduced space within the pores of
the MCM-41 material and the lower amount of surface hydroxyl groups [65,
66]. Dharman et al. further investigated the influence of the structure of the silica
support on the microwave-assisted synthesis of allyl glycidyl carbonate. It was
found that silica supports with relatively large pores (>4 nm) showed higher
catalytic activities and that microwave induction leads to a dramatically decrease
of the reaction time [67]. Sakai et al. could show that the acidity of the surface
hydroxy groups also affects the catalytic activity of the SILP catalysts.
Thereby, immobilization of phosphonium halides on Toyonite and especially
basic alumina led to a lower catalytic activity of the SILP material, than the use of
silica as the support material [62].
As mentioned before, the use of hydroxyl-functionalized ILs as catalysts results
in a higher reactivity. This also applies to SILP catalysts with OH-containing ILs,
which was shown by the comparison of immobilized HOETABr and
unfunctionalized ETABr as catalyst [73].
Another possibility for the synthesis of organocatalytic active SILPs is the
polymerization or copolymerization of vinyl-functionalized ILs (Scheme 1.6).
The first example for this kind of SILP catalyst was given by Han and coworkers
who polymerized 3-butyl-1-vinylimidazolium chloride (VBImCl) or copolymerized
the IL and divinylbenzene (DVB) with AIBN as the initiator (Scheme 1.6, a).
The resulting insoluble highly cross-linked polymer-supported IL was able to
catalyze the formation of PC with a yield of 97 % at 110  C (60 bar CO2, 7 h,
0.2 mmol active sites). In comparison to BMImCl, it also showed slightly higher
reactivity and was easily recyclable for at least four times by simple filtration [74].
Recently, Wang and Cheng prepared a series of hydroxyl- and carboxyl-functionalized
polymers and copolymers of VImBr with DVB (Scheme 1.6, b). The quantitative
conversion of PO to PC was achieved after 4 h at 130  C and 25 bar CO2 pressure
with 1.0 mol-% of the immobilized CC2ImBr catalyst. The superior reactivity of
the functionalized catalyst derives again from the activation of the epoxide through
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 15

NR3, PR3
or
NR3+ Cl- PR3+ Cl-

HN N
Cl RX
CMPS N
N X-
N N+ R

N N
RX X-
N N N
H H N+ R
PDVB-Im

Scheme 1.7 SILP synthesis with functionalized PS resin as the solid phase

hydrogen bonds inducing a positive effect on the ring-opening [75]. Xiong


et al. synthesized cross-linked polymeric nanoparticles (CLPN) as SILP catalysts
for the synthesis of various carbonates including PC. The copolymerization of
4-vinylbenzyl-triphenylphosphorous chloride or the respective butyl analog and
ethylene glycol dimethacrylate (EGDMA) in a ratio of 2:1 initiated by AIBN gave
the most active catalysts (Scheme 1.6, c). The SILP materials could catalyze the
formation of PC with excellent yields of >99 % and a selectivity of up to 100 %
(30 bar CO2, 3 h, CLPN-PBu3Cl: 140  C, PPh3Cl: 160  C) and were reusable for at
least 4 times [76, 77]. The same group applied this method to develop a
3-(2-methoxy-2-oxyl ethyl)-imidazolium chloride (VCC2MImCl)-based CLPN
catalyst (EGDMA/IL/AIBN: 2/1/~0.07) (Scheme 1.6, d ). The cycloaddition of
CO2 to PO was conducted with nearly quantitative yield and 100 % selectivity at
relatively harsh conditions (160  C, 50 bar CO2, 12 h), when compared to the
phosphorous-based CLNPs [78]. In all of these cases, the copolymers showed
higher reactivity than the polymerized vinyl-substituted ILs. This derives most
probably due to the very high density of active sites in the homopolymers, leading
to the insufficient use of catalytic active centers [75].
Furthermore, the grafting of ILs on functionalized polymers displays an alter-
native method for the preparation of SILP catalysts. Through the coupling of
tertiary amines or phosphines with chloromethylated polystyrene resin (CMPS,
cross-linked by copolymerization with varying degrees of DVB; DC, degree of
cross-linking), it is possible to synthesize polymer-supported quaternary onium
salts. For the immobilization of imidazolium salts on CMPS, imidazole firstly is
reacted with the resin and is functionalized in the second step. As an alternative,
commercially available cross-linked divinylbenzene polymer, grafted with
1-(3-aminopropyl) imidazole (PDVB-Im), is converted to the corresponding halide
to obtain the desired product (Scheme 1.7). The PS-based SILP catalysts with the
corresponding supporting materials and reaction conditions for the cycloaddition
reaction of CO2 are given in Table 1.2.
Immobilized quaternary onium salts and their application as SILP catalysts for
the cycloaddition of CO2 to PMO were first introduced in the year 1993. It was
pointed out that the length of the alkyl chain attached to the supported IL strongly
affects the reactivity, which reached a maximum for R ¼ n  Bu [79].
16

Table 1.2 Catalytic formation of CCs through SILP catalysts obtained by grafting of ILs on functionalized PS resins
Supported IL DC [%] Substrate T [ C] t [h] pCO2 [bar] Cat. loading [mol-%] Conversion [%] Selectivity [%] Ref.
P(nBu)3Cl 2 PMO 100 24 atma 1.0 93 100 [79]
N(nBu)3Cl 2 GMA 110 6 atmb 1.0 * * [80]
c
N(Me)3Cl 1 PO 100 24 80.0 5.0 99 100 [81]
N(Me)3Cl 2 PO 150 5 20.0 1.1 96 99 [82]
NEt(C2H4OH)2Brd 2 PO 110 4 20.0 2.0 99 99 [83]
P(nBu)3Cl 2 PO 150 6 50.0 3.2e 93 100 [84]
HEImI 6 PO 120 4 25.0 1.6 99 99 [85]
HeMImCl 1 AGE 120 4 15.5 11.0e 83 99 [86]
DHPImBr 2 PO 125 3 20.0 1.2 99 97 [87]
HEImBrf 1 PO 140 4 20.0 0.44 98 100 [88]
CC3ImBrf 1 PO 140 4 20.0 0.44 96 100 [89]
*k-values are given
PMO 2-phenoxymethyloxirane, GMA glycidyl methacrylate, HE 1-(2-hydroxyethyl), HeM 1-n-hexyl-3-methyl, DHP 1-(2,3-di-hydroxyl-propyl)
a
CO2 flow, 40 mL/min
b
10 mL/min
c
Ion exchange resin
d
Prepared by quaternization of PS-N(C2H4OH)2 with EtBr
e
wt.-%
f
Resin, PDVB-Im
M.H. Anthofer et al.
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 17

This indicates that more sterically hindered alkyl structures inhibit the approach
of the epoxide to the reactive sites, while shorter chain lengths of the ILs result in a
less lipophilic catalyst [80]. A series of imidazolium-based SILP catalysts were
prepared later. The ILs containing hydroxyl groups showed higher catalytic activity
than the unfunctionalized ILs due to the epoxide activation through hydrogen bonds
[85]. This was also shown by other groups through comparison of the catalytic
activity of PS-HEImBr, PS-CC3ImBr, and unfunctionalized ILs under the investi-
gated reaction conditions [88, 89]. In addition to that, ILs containing two or more
hydroxyl groups were immobilized to further increase the activity of the resulting
SILP materials [83, 87]. The reactivity of the PS-HEImX also strongly depends on
the anion and decreased in the order of I ~ Br > Cl in accordance to the
nucleophilicity [85].
The quaternization of cross-linked (2 %) poly-(4-vinylpyridine) with halides is
another example for the preparation of polymer-bound IL catalysts. The resulting
pyridinium-based IL (best catalyst: RX ¼ 1-n-butylbromide) is able to catalyze the
fixation of CO2 to glycidyl phenyl ether in a yield of >99 % (100  C, 1 bar CO2,
12 h, 5.0 wt.-% cat. loading) [90]. The protonation of polyanilines with HX also
creates highly active SILP catalysts. When HI was used as the doping agent
(protonation degree: 3.0 mmol g1), the resulting SILP-like material could catalyze
the formation of PC at 115  C in a yield of 99 % (50 bar CO2, 6 h, 20 wt.-%)
[91]. Han and coworkers prepared porous polymer bead SILP catalysts by suspen-
sion copolymerization of 1-vinylimidazole and DVB with AIBN as initiator and
subsequent quaternization of the supported imidazole. The 3-n-butyl-iodide-
functionalized IL, supported on a polymer with a high-surface area, and amount
of immobilized IL, gave more active catalysts for the conversion of AGE, whereas
nonporous polymer beads as carrier material resulted in a decrease in reactivity
[92]. Tubular microporous organic networks, with imidazolium iodide as the active
species for the coupling of CO2 with epoxides, are also accessible via Sonogashira
coupling of a tetrahedral building block and the corresponding diiodoimidazolium
salt. The generated SILP catalyst gave a yield of 87 % PC using a relatively low
catalyst loading of 0.065 mol-% (150  C, 10 bar CO2, 10 h) [93]. Alternatively,
tubular heterogeneous SILP catalysts can be obtained by esterification of carboxy
group containing carbon nanotubes with hydroxyl-functionalized imidazolium
salts. The grafting of MImI resulted in a SILP which catalyzed the formation of
SC in a yield of 97 % at 115  C (6 h, 18.2 bar CO2, 2.1 wt.-%) [40]. Recently, He
and coworkers developed a SILP, which performs as a homogeneous catalyst under
supercritical conditions but is easily separable after the completion of the reaction.
This was achieved by the formation of a fluoroacrylate/chloromethylstyrene copol-
ymer (10/1) and subsequent reaction with P(nBu)3 [94]. The preparation of multi-
layered SILP catalyst illustrates a combined method for the immobilization of
imidazolium-based IL on silica or SBA materials and organic polymers. For that
purpose, a bis-vinylimidazolium salt is linked to a mercapto-modified polystyrene
or silica (amorphous or SBA-15) by the thiol-ene reaction. High excess of the IL
results in a high loading of imidazolium units within the SILP material (up to
3.40 mmol g1). SBA-15-supported imidazolium bromides showed the highest
18 M.H. Anthofer et al.

NR3, PR3, RIm, guanidine


HO PEG OH Br PEG Br Br- +IL PEG IL+ Br-

Scheme 1.8 Preparation of PEG-supported ILs

catalytic activity in the high-throughput conversion of PO (84 %, 95 % selectivity)


under scCO2 (150  C, 100 bar CO2, 1 mol-%, 3 h) because of the higher surface area
of the silica material [95]. All of the above listed insoluble SILP catalysts based on
PS or other polymers are recyclable through simple filtration from the reaction
mixture and therefore provide the capability of an effective chemical fixation of
CO2 to cyclic carbonates.
As an inexpensive, thermally stable, and nontoxic supporting material, polyethylene
glycols (PEGs) represent a green alternative to the already mentioned supporting
materials. The easiest and most applied method for the synthesis of PEG-supported
ILs displays the bromination of the polymer chain ends and the subsequent reaction
with the desired IL (Scheme 1.8).
The first example of the application of this class of SILP catalysts for the
synthesis of cyclic carbonates was shown by He et al. in 2006. The coupling of
NBu3 with PEG6000 resulted in an increase of catalytic activity due to the changes of
physical properties of the reaction mixture. Under scCO2 (120  C, 80 bar CO2), an
excellent PC yield (98 %) and selectivity (99 %) could be achieved after 6 h
reaction time with a relatively low catalyst loading (0.5 mol-%). Moreover, the
catalyst could be recycled by simple filtration from the cooled reaction mixture
[96]. The reaction of a pentaalkylguanidine with PEG6000 was realized by the same
group. The prepared catalysts gave an even better PC yield (99 %) at milder
reaction conditions (110  C, 40 bar CO2, 4 h, 0.5 mol-%). However, the catalyst
recycling had to be carried out by extraction with diethyl ether [97]. In addition to
that, He and coworkers also synthesized a series of PEG150-grafted imidazolium
salts (MMImX, MImBr, X ¼ Cl, Br, I) and guanidinium bases (DBN, DBU, TBD).
The resulting TBD-based catalyst was able to catalyze the formation of PC in nearly
quantitative yield (>99 %) at relatively mild reaction conditions (120  C, 10 bar
CO2, 8 h, 1 mol-%). Most probably, the formation of an activated carbamate
species, through the reaction of CO2 with the secondary amino group of TBD,
which was detected by in situ FTIR, enhances the reactivity [98]. Furthermore,
these catalysts can also be used for the catalytic conversion of CO2 with aziridines
to form oxazolidinones. Hereby, the DBN-based catalysts showed not only high
reactivity but also good regioselectivities [99]. An example for a phosphonium
bromide-based SILP catalyst was given by Bhanage et al. through the reaction of
PPh3 with PEG600. The conversion of SO or epichlorohydrin to the respective
carbonates could be catalyzed in excellent yields and selectivities (>98 %),
whereas the formation of PC was rather poor (60 % yield) under the investigated
reaction conditions (100  C, 27 bar CO2, 3 h, 5 wt.-%) [100]. Very recently,
Park and coworkers published their work on the conversion of AGE with CO2,
catalyzed by imidazolium salts supported on PEG4000. For the grafting of the IL,
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 19

OH
X IL+ X-
O O IL = +NR3, MEIm; X = Cl, Br, I
HO O
OH NH n
O O + -
O IL X
HO
NH2 n OH
RX O R = Me, Et, n-Pr, n-Bu; X = Cl, Br, I
O
HO O
+
NR3 n
X-

Scheme 1.9 Synthesis of chitosan-based SILP catalysts

one chain-end hydroxyl group of the polymer is firstly transformed with succinic
anhydride to the corresponding succinic acid and then esterified with HAImX
(A ¼ Me, Et, n-Bu; X ¼ Cl, Br, I). The catalyst with immobilized MImCl showed
the highest reactivity reaching a conversion and selectivity of 98 % at 140  C
(18.9 bar CO2, 4 h, 2.3 mol-%) [101].
Considering the biocompatibility, biodegradability, and high availability of the
biopolymer chitosan (CS), it represents a suitable supporting material for ILs in
terms of green chemistry. The hitherto known methods for the grafting of ILs on
chitosan are shown in Scheme 1.9.
The immobilization of 3-chloro-2-hydroxypropyl trialkylammonium halides on
chitosan was the first CS-SILP catalyst investigated for the cycloaddition of CO2
with epoxides. He and coworkers have shown that CS as a supporting material
could enhance the catalytic activity of the immobilized IL through a synergistic
effect, which derives from the coexistence of hydrogen bond donors (-OH of CS)
and the halide anions of the ammonium moiety (N+R3X) [102, 103]. With
CS-N+Me3Cl as the model catalyst (1.7 mmol-%), an excellent PC yield (98 %)
and selectivity (>99 %) were reached, however under relatively harsh reaction
conditions (160  C, 40 bar CO2, 6 h) [102]. Park et al. synthesized a series of
CS-SILP catalysts through the quaternization of the CS-amino group with alkyl
halides [103, 104]. The efficiency of the catalyst preparation could be increased
drastically through microwave irradiation, yielding CS-N+Me3I as the most
reactive species. The resulting system was able to catalyze the formation of PC
with good yields (89 %) and excellent selectivities (>99 %) at comparably mild
reaction conditions (120  C, 11.7 bar CO2, 6 h, 1.6 mmol-%) [104]. Moreover, SILP
catalysts are accessible through the grafting of Br-EMImBr on CS. In comparison to
the PS-bound catalysts, the CS-EMImBr showed higher activity due to the already
mentioned synergistic effect of the carrier material. As a result, excellent yields and
selectivities (96 % and >99 % for PC) could be achieved (120  C, 20 bar CO2, 4 h,
1 mmol-%) [105]. Another example for a biopolymer as carrier material was given
by Park et al. who used carboxymethyl cellulose for the non-covalently immobili-
zation of imidazolium salts. DFT calculations have shown that the nucleophilic
attack at the epoxide is facilitated via hydrogen bonds from the carboxyl and
hydroxyl functional groups of the biopolymer. This synergistic effect results in a
20 M.H. Anthofer et al.

high activity of the iodide-containing SILP catalyst for the formation of PC (yield
and selectivity >98 %) under relatively mild conditions (110  C, 18 bar CO2, 2 h,
1.2 mmol-%) [106].

1.4 Formation of Cyclic Carbonates Catalyzed


by Nitrogen Donor Bases

Organic nitrogen bases represent another class of metal-free catalysts for the
transformation of carbon dioxide. The most commonly used bases for this purpose
are illustrated in Fig. 1.6.
To the best of the authors’ knowledge, the first example for the nitrogen
base-mediated cycloaddition of CO2 to epoxides was reported by Sartori et al. in
2003. The guanidine MTBD (7-methyl-1,5,7-triazabicyclo[4.4.0]dec-5-ene) and
the supported analog (MCM-41 as carrier material) were used as catalysts for
the formation of a variety of carbonates including SC. However, harsh reaction
conditions (140  C, 50 bar CO2, 4 mol-%) and long reaction times (20–40 h) were
necessary to obtain full conversion of SO (95 % yield) [107]. Furthermore, TBD
was used as catalytic active species supported on silica or as molecular
organocatalyst in several other works [108–110]. The reactivity of the catalysts in
the cycloaddition reaction is strongly influenced by their basicity. Therefore, the
use of TBD, which represents a stronger base than, for example, DBU, resulted in
more active organocatalysts [109]. In addition, Sun and coworkers have shown that
the surface hydroxyl groups of the silica supporting material dramatically enhance
the catalytic activity of TBD. For this purpose, the reactivity of the TBD/SiO2
catalyst was compared with a methyl-modified silica species. It was pointed out that
the surface hydroxyl groups play an important role in the reaction, since the conversion
of PO dropped from >99 to <1 % (150  C, 20 bar CO2, 20 h, 6.4 wt.-%) when the
modified material was used as catalyst [108]. A possible application of the well-known
transesterification catalyst DMAP for the chemical fixation of CO2 with epoxides was
first discovered by Manikandan et al. in 2004. It was found that DMAP, which should
be used as the cocatalyst, was able to catalyze the conversion of epichlorohydrin to the

N
NH
N N
N N
N N N
N
R
DMAP TMG R = H; TBD DBU
R = Me; MTBD

Fig. 1.6 Structures of most frequently used bases for CO2 activation
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 21

respective carbonate by itself [111]. This was also shown by Jones and Shiels who used
homogeneous and SBA-15-supported DMAP for the formation of PC. The molecular
DMAP showed slightly higher reactivity (85 % yield, 100 % selectivity) under the
investigated reaction conditions (120  C, 17.2 bar CO2, 4 h, 0.4 mol-%) [112]. Another
class of N-donor base catalysts for the cycloaddition of CO2 were prepared by
Ratnasamy and coworkers via the covalent immobilization of adenine, guanine,
imidazole, and primary alkyl amines on SBA-15 [113, 114]. The resulting adenine
catalyst, which showed the highest reactivity, could catalyze the conversion of SO in
good yields (86 %) and excellent selectivities for SC (97 %) under comparably mild
reaction conditions (120  C, 6.9 bar CO2, 8 h) [114]. The non-covalently immobiliza-
tion of polyvinyl pyridine on silica also leads to organocatalytic active solid basic
materials. The optimized catalyst (10 wt.-% PVP on SiO2) gave 93 % yield of PC at
150  C and 55 bar CO2 after 6 h [115]. The recently reported preparation of TMG-SiO2
nanoparticles is another possibility for the synthesis of solid base catalysts. The
grafting of TMG on 3-chloropropyl-modified silica from rice husk ash gave amorphous
spherical particles with a mean size of 19.6 nm. The cycloaddition of CO2 to PO was
carried out with 92 % conversion and 98 % selectivity for PC at 130  C and 50 bar CO2
for 8 h [116].
Park and coworkers synthesized and applied melamine-bridged periodic
mesoporous organosilica (PMO) as organocatalyst for the cycloaddition of CO2
to epoxides. In PMO, the organic moieties are integrated in the silica framework
forming an ordered mesoporous type of organic–inorganic material. Through the
incorporation of large-sized melamine-based triorganosilsesquioxane as organic
linker, a well-ordered hexagonal mesostructural organocatalyst could be synthe-
sized. After 10 h at 100  C and 5.5 bar CO2, around 40 % conversion of PO to PC
with a high selectivity could be achieved [117]. The use of as-synthesized covalent
triazine frameworks (CTF) for the metal-free catalytic valorization of CO2 to
carbonates was reported by Roeser and coworkers. Under investigated reaction
conditions, the crystalline CTF was significantly less active than the amorphous
high-surface area (HSA) analog. Furthermore, the pyridinic CTF (CTF-P) showed
higher reactivity than the benzylic CTF, because of the higher basicity leading to an
increased extent of CO2 absorption. The CTF-P-HSA was able to catalyze the full
conversion of epichlorohydrin to the respective cyclic carbonate with a selectivity
of 96 % within 4 h under relatively mild reaction conditions (130  C, 6.9 bar CO2)
[118]. A possible mechanism for the formation of cyclic carbonates via nitrogen
base-catalyzed cycloaddition reaction of CO2 with epoxides is shown in
Scheme 1.10 with SiO2-TMG as the model catalyst [110, 116].
In the first step, CO2 is absorbed through reaction with the Lewis basic sites to
form a carbamate species [107, 110, 114, 116, 117]. The resonance stabilization
of the resulting cation enhances the stability of the cationic intermediate. The
nucleophilic attack of the carbamate anion at the less sterically hindered carbon
atom of the epoxide leads to the ring-opening of the epoxide. The formed anion
intramolecularly attacks the carbonyl carbon atom forming a five-membered
cyclic species. Finally, PC is eliminated from the surface regenerating the
catalyst [110, 116].
22 M.H. Anthofer et al.

O
O
O N CO2
SiO2 N N

N N
SiO2 -
N N SiO2 N N
O
O O O O-
O

N
N
SiO2 N N
SiO2 N N
O O
-
O O O-

Scheme 1.10 TMG-SiO2-catalyzed cycloaddition reaction of CO2 to PO [116]

1.4.1 Binary Catalyst Systems for the Cycloaddition


of Epoxides and CO2

In the following chapter, binary catalyst systems for the chemical fixation of CO2 and
epoxides (Scheme 1.1) are introduced. All of them show a cooperative effect between
a hydrogen bond donor and a nucleophile and the mechanism is proposed to be
identical according to HFILs (Scheme 1.2). The first approaches of binary catalyst
systems were carried out using phenol as hydrogen bond donor [119–121]. Neverthe-
less, using phenol as catalyst results in a low selectivity because also the side reaction
to 1-phenoxypropan-2-ol takes place. Therefore, different groups did a broad catalyst
screening with different aliphatic and aromatic alcohols and found 1,2-benzenediol
(BDO) and 1,2,3-benzenetriol (BTO) in combination with TBAB or TBAI as
cocatalyst to be the most active [51, 121]. Furthermore, there are other catalytic
systems based on cellulose [122], β-cyclodextrin [123], lecithin [124], pentaerythritol
[125], and lignin [52] in combination with potassium iodide (KI) as cocatalyst.
An overview of these binary catalyst systems is listed in Table 1.3 with their optimum
reaction conditions.
Other reports describe the influence of the hydrogen bond donor character using
theoretical calculation methods. It was found that the catalytic activity strongly
depends on the distance between the two hydroxyl groups [121, 126]. Additional
calculations concerning the BTO/TBAI system suggest a stabilization of the
Table 1.3 Overview of selected effective binary catalyst systems
Catalyst Co-catalyst Molar ratioa Substrate T [ C] p(CO2) [bar] t [h] Conv. [%] Sel. [%] Recycling runs Reference
b
Cellulose KI 66 /1.5 PO 110 20 2 99 99 5 [122]
β-cyclo-dextrin KI 8.7b/2.5 PO 120 60 4 98 99 5 [123]
Lecithin KI 1.25/1.25 PO 100 20 4 98 98 – [124]
Lignin KI 0.67c/2 PO 140 20 12 91 99 5 [52]
BDO TBAB 5/1 PO 100 30 0.5 92 99 – [121]
BTO TBAI 5/5 PO 45 10 16 100 99 2d [51]
a
In relation to the substrate (mol-% cat./mol-% co-cat.)
b
Amount in wt.-%
c
Mol-% of active OH-groups
d
Loss of activity with each run
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts
23
24 M.H. Anthofer et al.

Scheme 1.11 Stabilization


of the oxyanion O
O O CO2
intermediate by BTO
O H O
H O H O
H
H
O O
H O
I
I

a O
+2 eq. CH3OH O
O OH
CO2 O O HO
O O
R1 R1
R1

b O
CO2 2 CH3OH H2O
O O

Scheme 1.12 Synthesis of DMC by (a) transesterification of cyclic carbonates with MeOH or
(b) by direct synthesis of CO2 and MeOH

oxyanion before and after the insertion of CO2 by the adjacent hydroxyl groups of
BTO (see Scheme 1.11). After activating the epoxide via hydrogen bonding, the
following oxyanion intermediate is stabilized by hydrogen bonding of the neigh-
boring hydroxyl groups. For the same reason, glycerol is also an efficient catalyst
for the chemical fixation of CO2 with epoxides [82]. However, glycerol is not a
suitable catalyst since during the reaction glycerol carbonate (GC) is also formed.
This prevents the systems to be recycled for several times since glycerol is also
consumed as substrate. Nevertheless, GC offers a lot of other opportunities and is an
attractive target molecule for the chemical fixation of CO2 [127].

1.5 Synthesis of Dimethyl Carbonate (DMC)

Utilizing CO2 as a feedstock for the green synthesis of DMC, there are two possible
routes for reaching this goal in an environmentally friendly and cheap way. DMC
can be synthesized either by the transesterification of a cyclic carbonate with
methanol (MeOH) (Scheme 1.12a) or by the direct synthesis from CO2 and
MeOH (Scheme 1.12b). The first route represents an indirect way for DMC
synthesis out of CO2 because cyclic carbonates can be produced from CO2 and
epoxide, as described in Sects. 1.2, 1.3, and 1.4. High CO2 pressure in the
transesterification reaction of a cyclic carbonate with MeOH is required to prevent
decarboxylation of the cyclic carbonate and is not directly involved in the reaction
procedure. For this reason, only the one-pot synthesis out of an epoxide, CO2, and
MeOH is mentioned in the following sections of this chapter.
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 25

n=1, X=Br: Me2NC6NMe3Br


n=5, X=Br: Me2NC6NMe2HexBr
n=3, X=I: Me2NC6NMe2BuI
N n=3, X=NTf2: Me2NC6NMe2BuNTf2
N n
X n=3, X=Br: Me2NC6NMe2BuBr
n=3, X=Cl: Me2NC6NMe2BuCl
n=3, X=BF4: Me2NC6NMe2BuBF4

Scheme 1.13 Overview of ILs bearing a tertiary amino moiety and a quaternary ammonium group

1.5.1 One-Pot Synthesis of DMC from Methanol,


CO2 and Epoxides

Wang et al. showed a good activity toward the one-pot synthesis out of methanol, CO2,
and EO for using the mixture KI:K2CO3 (weight ratio 1:1) under supercritical condi-
tions (150  C, 150 bar CO2 pressure, 2 h, 2 wt.-% catalyst) [128]. A detailed investi-
gation showed an increasing yield of the desired products at higher temperature,
whereas the CO2 pressure has more or less no effect on the yield. Increasing the EO
content results in a better selectivity. Kinetic studies confirmed a relatively high reaction
rate for EC formation [129]. Due to a low reaction rate for the transesterification with
MeOH to form DMC, this step is proposed to be the rate-determining step in the one-pot
synthesis. Also, amines were found to be active and therefore DBU served as a suitable
catalyst for the reaction of glycidyl phenyl ether, MeOH, and CO2 (150 bar) at 150  C to
reach high yields of DMC and the corresponding 1,2-diol (>95 %) [130]. Extension of
this one-pot coproduction of DMC and 1,2-diols to the synthesis of other symmetric
dialkyl carbonates leads to yields of 63 % diethyl carbonate (DEC) and 44 % dibutyl
carbonate (DBC) under similar reaction conditions. Changwen et al. adopted a mixture
of BMImBF4 and sodium methanolate (NaOMe) as catalyst for the one-pot synthesis
[131]. Under optimized reaction conditions (40 bar, 5 h, 150  C), the conversion of PO
was >95 %, reaching a yield of 67 % DMC paired with a selectivity <10 % for the
by-products. A binary catalyst system consisting of TBAB as the cycloaddition catalyst
to form cyclic carbonates and tri(n-butyl)amine (NBu3) as a possible transesterification
catalyst was introduced [132]. Under optimized reaction conditions (2.5 mol-% catalyst
loading, 150  C, 150 bar, 8 h), PO, SO, and 2-(phenoxymethyl)oxirane (PMO) exhibit
59 %, 80 %, and 84 % of DMC yield, respectively.
An inorganic base (K2CO3) and a phosphonium halide-functionalized PEG
(BrBu3PPEG6000PBu3Br) binary system was developed as heterogeneous catalyst
and shown to be active for the one-pot synthesis of DMC even under low CO2
pressure (2 bar) [133]. The cycloaddition to form the cyclic carbonate was carried
out at 120  C under 10 bar CO2 for 6 h. Afterward, MeOH was added and the mixture
was further heated for 1.5 h at 100  C while removing DMC via distillation. It was
possible to reach yields up to 97 % DMC even in the fourth run of the catalytic cycle.
Moreover, leaching of the catalyst could be also excluded via 31P-NMR after recov-
ering the catalyst by filtration. Novel ILs bearing both a tertiary amino moiety and a
quaternary ammonium group were synthesized and tested toward the one-step syn-
thesis of DMC (Scheme 1.13) [134]. Under 20 bar CO2 pressure at 150  C and a
26 M.H. Anthofer et al.

catalyst loading of 1.3 mol-%, quantitative conversion of EO was observed after 8 h


with a selectivity of 74 % toward DMC using Me2NC6NMe2BuI as catalyst. Addi-
tionally, the catalyst could be reused for several times.
Zheng and coworkers systematically screened binary catalysts for the one-pot
synthesis of DMC and found that in presence of H2O, commercial quaternary
ammonium salts/K2CO3, pyridinium salts/K2CO3, and KI/K2CO3 were remarkably
effective for DMC synthesis with no need for separation of EC [135].
Nearly quantitative conversion of EO and 82 % yield of DMC were obtained under
relatively mild reaction conditions. The cycloaddition was carried out at 120  C under
25 bar CO2 pressure for 1 h with 0.018 mol-% H2O, 0.45 mol-% K2CO3, and
1.88 mol-% KI. After adding MeOH, the temperature was decreased to 60  C and
further reacted for 0.5 h. This system could be easily recovered and reused for
12 times without obvious loss of its catalytic activity. However, this study shows
an incompatibility of imidazolium salts with inorganic alkaline toward the one-pot
synthesis of DMC possibly because of the reactive C2 proton of the imidazolium ring.
Newer research effort showed that also a CTF is an active catalyst for synthesizing
DMC out of PO, CO2, and MeOH in one-pot [118]. At 20 bar CO2 and 160  C, a
conversion of 95 % PO was observed with a selectivity of 17 % toward DMC. This is
the first example using a porous organic coordination polymer for DMC synthesis.

1.5.2 Direct Synthesis of DMC from CO2 and Methanol

A first example for the direct synthesis of DMC was carried out using different
bases and methyl iodide (MeI) as a promoter for the reaction [136], where dimethyl
ether (DME) is formed as a by-product. Further investigations for this reaction
showed the highest activity for K2CO3 as a catalyst using MeI as the promoter for
the reaction [137]. Arai et al. observed two maxima in DMC formation at
45 and 80 bar CO2 pressure while DME formation decreases with higher pressure.
Additional mechanistic studies suggest a parallel pathway for DMC and DME
formation, whereas MeI is involved in both. Using other alcohols than MeOH
leads to a lower reactivity. In both cases, the optimized reaction conditions lead
to a yield of less or about 4 % DMC with a rather bad selectivity. Catalytic
experiments with potassium methanolate (KOMe) as catalyst and MeI as promoter
raised the yield of DMC to 16 % with a selectivity of 99 % at 80  C and 20 bar
CO2 pressure, whereas there is a second maximum of yield at 72 bar [138]. Thus,
higher pressures are not required for an effective synthesis of DMC. Wang
et al. used KOH as a catalyst for the direct synthesis of DMC with MeI and studied
the effects of adding the IL EMImBr to the reaction mixture [139]. Therefore, an
increase of the yield from 8 % DMC without EMImBr to 11 % in the presence of the
IL was observed at a CO2 pressure of 20 bar. The reason for higher DMC yield in
presence of EMImBr may be ascribed to the strong polarity and electrostatic field of
the IL, which may stabilize the charged intermediate. Interestingly, also in this case,
two maxima for the yield of DMC were found at 20 and 73 bar CO2 pressure, which
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 27

is similar to previous described works. Later work deals with the thermodynamics
of DMC synthesis and it could be shown that the reaction does not spontaneously
occur in view of thermodynamics using K2CO3, KOMe, or KOH as catalyst [140].
This may offer new possibilities for the design of effective organocatalytic systems
for the synthesis of DMC.

1.6 Further Organocatalytic Transformations of CO2

1.6.1 Organocatalytic Reduction of CO2


into Value-Added Chemicals

Employing CO2 as a true C1 building block and therefore preparing a wide spectrum
of chemicals, catalytic reactions that are able to promote the complete deoxygenation
of CO2 are required. For example, NHCs reversibly form zwitterionic CO2 adducts,
which are considered to be key intermediates in the reductive deoxygenation of CO2.
The role of NHC during catalysis was investigated theoretically and experimentally
[141, 142], and for using diphenylsilane as a sacrificial reducing agent, CH3OH could
be detected upon workup.
Cantat and coworkers developed the organocatalytic formylation reaction of
N–H bonds by using CO2 and hydrosilanes to yield formamides. To achieve
complete deoxygenation, they proposed that the amide function could react in a
cascade reaction with a nucleophile (Scheme 1.14), such as an amine, which results
in a substitution of the C═O bond in the formamide derivative [143–145].
The formylation was successfully catalyzed by nitrogen bases, such as TBD with
using silanes instead of H2 as reducing agent, allowing the utilization of an
organocatalyst which avoids any toxic and/or expensive metal. The reaction takes
place at 100  C under CO2 pressure <3 bar without an additional solvent and
promotes the formylation of a wide spectrum of amines due to the mild reaction
conditions, while using H2 as reducing agent, the reaction is limited to only a few
amines [143]. Nevertheless, this system is limited to basic amines and only

Scheme 1.14 Principles


of complete deoxygenation
of CO2 and the cascade
reaction for substitution of
C═O bond in the
formamide derivative
28 M.H. Anthofer et al.

a
H TMHP H ΔT H
H H OH
ΔT O
TMHP H H B(C6F5)3 H
b H +E
O t3 S
+ iH
O B(C6F5)3 B(C O ΔT
6F
5)
3 CH4
H O SiEt3
TMP =
NH

Scheme 1.15 Formation of (a) methanol and (b) methane out of the boratocarbamate-TMPH
ion pair

proceeds at 100  C, whereas low temperatures and broad application spectra are
required for CO2 recycling in an environmentally friendly way. Further investigations
demonstrated that the formylation step is also effectively catalyzed by NHCs.
NHCs are efficient catalysts for the conversion of amines, anilines, imines, and
N-heterocycles at room temperature, and, therefore, NHCs are more attractive for a
sustainable recycling of CO2 [144]. Also, the cascade reductive functionalization of
CO2 into benzimidazoles, quinazolinones, formamidines, and their derivatives is
catalyzed by NHCs and TBD at 70  C [145]. The results show a difference in
selectivity between the two classes of organocatalyst and suggest a different activation
pathway in the reductive functionalization of CO2, which has to be confirmed by
further investigations.
Another way for the nonmetal-mediated reduction of CO2 is the use of frustrated
Lewis pairs (FLPs) [146]. Therefore, the stoichiometric reduction of CO2 to
CH3OH was investigated in detail and it was found that a system consisting of
2,2,6,6-tetramethylpiperidine (TMP) and tris(pentafluorophenyl)borane [B(C6F5)3]
activates CO2 to form a boratocarbamate-TMPH ion pair (Scheme 1.15)
[147]. Heating this ion pair at 160  C under a H2 atmosphere leads quantitatively
to CH3OB(C6F5)2 after 6 d, and after vacuum distillation of the solvent, 17–25 %
CH3OH could be isolated. Another study dealing with an analogous FLP showed a
conversion of this species in the presence of triethylsilane (TES) to a silylcarbamate
and the known ion pair [TMPH]+[HB(C6F5)3], which was recently shown to react
with CO2 via transfer of the hydride from the hydridoborate to form the
formatoborate [TMPH]+[HC(O)OB(C6F5)3] [148]. In the presence of extra B
(C6F5)3 (0.1–1.0 equiv.) and excess TES, the formatoborate is rapidly hydrosilated
to form a formatosilane and regenerate [TMPH]+[HB(C6F5)3]. The formatosilane
in turn is rapidly hydrosilated by the B(C6F5)3/Et3SiH system to CH4, with
(Et3Si)2O as the by-product. At low concentration of Et3SiH, it is possible to isolate
intermediate CO2 reduction products, whereas addition of more CO2/Et3SiH results
in resumed hydrosilylation. All the results indicate a robust, living tandem catalytic
system for the deoxygenative reduction of CO2 to CH4.
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 29

1.7 Conclusions

It is shown that many promising concepts for the development of highly efficient
organocatalysts for the transformation of carbon dioxide to (bulk) chemicals are at
hand. Particularly in the field of the synthesis of cyclic carbonates, the variation of
organic substituents allows for tailoring the desired parameters in order to enhance
the catalyst activity in a similar fashion as it is possible with metal catalysts.
A drawback of most organocatalysts is still that usually rather high temperatures
and CO2 pressures are required. However, in the last years, significant improve-
ments could be achieved. Therefore, different catalysts were developed which
allow the activation of the epoxide and/or CO2 leading to lower reaction temper-
ature and CO2 pressure. Furthermore, the simple separation and reuse of a catalyst
for several times is very important for industrial application, which was proven and
realized for IL catalysts in the formation of cyclic carbonates. Alternative ways for
sustainable CO2 fixation represent the synthesis of DMC and the reduction to
methanol or methane. Nevertheless, the known organocatalytic systems still need
optimization regarding catalyst design to reach satisfying conversion under mild
reaction conditions. Considering all these facts, organocatalytic systems open the
possibility for a green and sustainable valorization of CO2 with regard to a CO2-
emission point of view as well as for industrial applications.

References

1. Brundtland GH (1987) United Nations report: our common future, Oxford University Press.
2. BP Statistical Review of World Energy (2012) Statistical review of world energy. BP p. l. c.
3. Aresta M, Dibenedetto A (2004) The contribution of the utilization option to reducing the CO2
atmospheric loading: research needed to overcome existing barriers for a full exploitation of the
potential of the CO2 use. Catal Today 98(4):455–462, http://dx.doi.org/10.1016/j.cattod.2004.
09.001
4. Bozell JJ, Petersen GR (2010) Technology development for the production of biobased
products from biorefinery carbohydrates-the US Department of Energy’s “Top 10” revisited.
Green Chem 12(4):539–554. doi:10.1039/B922014C
5. Marshall A-L, Alaimo PJ (2010) Useful products from complex starting materials: common
chemicals from biomass feedstocks. Chem Eur J 16(17):4970–4980. doi:10.1002/chem.
200903028
6. Ruppert AM, Weinberg K, Palkovits R (2012) Hydrogenolyse goes Bio: Von Kohlenhydraten
und Zuckeralkoholen zu Plattformchemikalien. Angew Chem 124(11):2614–2654. doi:10.1002/
ange.201105125
7. Arakawa H, Aresta M, Armor J, Barteau M, Beckman EJ, Bell AT, Bercaw JE, Creutz C,
Dinjus E, Dixon DA, Domen K, Dubois DL, Eckert J, Fujita E, Gibson DH, Goddard WA,
Goodman WD, Keller J, Kubas GJ, Kung HH, Lyons JE, Manzer L, Marks TJ, Morokuma K,
Nicholas KM, Periana R, Que L, Rostrup-Nielson J, Sachtler WMH, Schmidt LD, Sen A,
Somorjai GA, Stair PC, Stults BR, Tumas W (2001) Catalysis research of relevance to
carbon management: progress, challenges, and opportunities. Chem Rev 101(4):953–996.
doi:10.1021/cr000018s
8. Omae I (2006) Aspects of carbon dioxide utilization. Catal Today 115(1–4):33–52, http://dx.doi.
org/10.1016/j.cattod.2006.02.024
30 M.H. Anthofer et al.

9. Omae I (2012) Recent developments in carbon dioxide utilization for the production of
organic chemicals. Coordination Chem Rev 256(13–14):1384–1405, http://dx.doi.org/10.
1016/j.ccr.2012.03.017
10. Song C (2006) Global challenges and strategies for control, conversion and utilization
of CO2 for sustainable development involving energy, catalysis, adsorption and
chemical processing. Catal Today 115(1–4):2–32, http://dx.doi.org/10.1016/j.cattod.
2006.02.029
11. Centi G, Iaquaniello G, Perathoner S (2011) Can we afford to waste carbon dioxide? Carbon
dioxide as a valuable source of carbon for the production of light olefins. ChemSusChem
4(9):1265–1273. doi:10.1002/cssc.201100313
12. Peters M, Köhler B, Kuckshinrichs W, Leitner W, Markewitz P, Müller TE (2011) Chemical
technologies for exploiting and recycling carbon dioxide into the value chain. ChemSusChem
4(9):1216–1240. doi:10.1002/cssc.201000447
13. Quadrelli EA, Centi G, Duplan J-L, Perathoner S (2011) Carbon dioxide recycling: emerging
large-scale technologies with industrial potential. ChemSusChem 4(9):1194–1215. doi:10.1002/
cssc.201100473
14. Gallezot P (2012) Conversion of biomass to selected chemical products. Chem Soc Rev
41(4):1538–1558. doi:10.1039/C1CS15147A
15. Olah GA, Prakash GKS, Goeppert A (2011) Anthropogenic chemical carbon cycle for a
sustainable future. J Am Chem Soc 133(33):12881–12898. doi:10.1021/ja202642y
16. Aresta M, Dibenedetto A (2007) Utilisation of CO2 as a chemical feedstock: opportunities
and challenges. Dalton Trans 28:2975–2992. doi:10.1039/B700658F
17. Cokoja M, Bruckmeier C, Rieger B, Herrmann WA, Kühn FE (2011) Transformation of
carbon dioxide with homogeneous transition-metal catalysts: a molecular solution to a global
challenge? Angew Chem Int Ed 50(37):8510–8537. doi:10.1002/anie.201102010
18. Drees M, Cokoja M, Kühn FE (2012) Recycling CO2? Computational considerations
of the activation of CO2 with homogeneous transition metal catalysts. ChemCatChem
4(11):1703–1712. doi:10.1002/cctc.201200145
19. Coates GW, Moore DR (2004) Discrete metal-based catalysts for the copolymerization of
CO2 and epoxides: discovery, reactivity, optimization, and mechanism. Angew Chem Int Ed
43(48):6618–6639. doi:10.1002/anie.200460442
20. Sakakura T, Kohno K (2009) The synthesis of organic carbonates from carbon dioxide. Chem
Commun 11:1312–1330. doi:10.1039/B819997C
21. Darensbourg DJ (2010) Chemistry of carbon dioxide relevant to its utilization: a personal
perspective. Inorg Chem 49(23):10765–10780. doi:10.1021/ic101800d
22. Decortes A, Castilla AM, Kleij AW (2010) Salen-complex-mediated formation of cyclic
carbonates by cycloaddition of CO2 to epoxides. Angew Chem Int Ed 49(51):9822–9837.
doi:10.1002/ange.201002087
23. Kember MR, Buchard A, Williams CK (2011) Catalysts for CO2/epoxide copolymerisation.
Chem Commun 47(1):141–163. doi:10.1039/C0CC02207A
24. Klaus S, Lehenmeier MW, Anderson CE, Rieger B (2011) Recent advances in CO2/epoxide
copolymerization—new strategies and cooperative mechanisms. Coordination Chem Rev
255(13–14):1460–1479, http://dx.doi.org/10.1016/j.ccr.2010.12.002
25. Lu X-B, Darensbourg DJ (2012) Cobalt catalysts for the coupling of CO2 and epoxides
to provide polycarbonates and cyclic carbonates. Chem Soc Rev 41(4):1462–1484.
doi:10.1039/C1CS15142H
26. Shaikh A-AG, Sivaram S (1996) Organic carbonates. Chem Rev 96(3):951–976. doi:10.1021/
cr950067i
27. Calo V, Nacci A, Monopoli A, Fanizzi A (2002) Cyclic carbonate formation from carbon
dioxide and oxiranes in tetrabutylammonium halides as solvents and catalysts. Org Lett
4(15):2561–2563. doi:10.1021/ol026189w
28. Wang J-Q, Dong K, Cheng W-G, Sun J, Zhang S-J (2012) Insights into quaternary ammonium
salts-catalyzed fixation carbon dioxide with epoxides. Catal Sci Technol 2(7):1480–1484.
doi:10.1039/c2cy20103h
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 31

29. Peng J, Deng Y (2001) Cycloaddition of carbon dioxide to propylene oxide catalyzed by ionic
liquids. New J Chem 25(4):639–641. doi:10.1039/B008923K
30. Dharman MM, Choi H-J, Park S-W, Park D-W (2010) Microwave assisted synthesis of
cyclic carbonate using homogeneous and heterogeneous ionic liquid catalysts. Top Catal
53(7–10):462–469. doi:10.1007/s11244-010-9473-0
31. Kawanami H, Sasaki A, Matsui K, Ikushima Y (2003) A rapid and effective synthesis of
propylene carbonate using a supercritical CO2-ionic liquid system. Chem Commun
7:896–897. doi:10.1039/B212823C
32. Sun H, Zhang D (2007) Density functional theory study on the cycloaddition of carbon
dioxide with propylene oxide catalyzed by alkylmethylimidazolium chlorine ionic liquids.
J Phys Chem A 111(32):8036–8043. doi:10.1021/jp073873p
33. Kim HS, Kim JJ, Kim H, Jang HG (2003) Imidazolium zinc tetrahalide-catalyzed coupling
reaction of CO2 and ethylene oxide or propylene oxide. J Catal 220(1):44–46, http://dx.doi.org/
10.1016/S0021-9517(03)00238-0
34. Wong W-L, Chan P-H, Zhou Z-Y, Lee K-H, Cheung K-C, Wong K-Y (2008) A robust ionic
liquid as reaction medium and efficient organocatalyst for carbon dioxide fixation.
ChemSusChem 1(1–2):67–70. doi:10.1002/cssc.200700097
35. Sun J, Zhang S, Cheng W, Ren J (2008) Hydroxyl-functionalized ionic liquid: a
novel efficient catalyst for chemical fixation of CO2 to cyclic carbonate. Tetrahedron Lett
49(22):3588–3591. doi:10.1016/j.tetlet.2008.04.022
36. Zhou Y, Hu S, Ma X, Liang S, Jiang T, Han B (2008) Synthesis of cyclic carbonates
from carbon dioxide and epoxides over betaine-based catalysts. J Mol Catal A-Chem
284(1–2):52–57. doi:10.1016/j.molcata.2008.01.010
37. Tsutsumi Y, Yamakawa K, Yoshida M, Ema T, Sakai T (2010) Bifunctional organocatalyst
for activation of carbon dioxide and epoxide to produce cyclic carbonate: betaine as a new
catalytic motif. Org Lett 12(24):5728–5731. doi:10.1021/ol102539x
38. Qi C, Jiang H (2010) Histidine-catalyzed synthesis of cyclic carbonates in supercritical
carbon dioxide. Sci China-Chem 53(7):1566–1570. doi:10.1007/s11426-010-4019-7
39. Wu F, Dou X-Y, He L-N, Miao C-X (2010) Natural amino acid-based ionic liquids as
efficient catalysts for the synthesis of cyclic carbonates from CO2 and epoxides under
solvent-free conditions. Lett Org Chem 7(1):73–78
40. Gong Q, Luo H, Cao J, Shang Y, Zhang H, Wang W, Zhou X (2012) Synthesis of cyclic
carbonate from carbon dioxide and epoxide using amino acid ionic liquid under 1 atm
pressure. Aust J Chem 65(4):381–386. doi:10.1071/ch11462
41. Zhou H, Zhang W-Z, Liu C-H, Qu J-P, Lu X-B (2008) CO2 adducts of N-heterocyclic
carbenes: thermal stability and catalytic activity toward the coupling of CO2 with epoxides.
J Org Chem 73(20):8039–8044. doi:10.1021/jo801457r
42. Kayaki Y, Yamamoto M, Ikariya T (2009) N-heterocyclic carbenes as efficient organocatalysts for
CO2 fixation reactions. Angew Chem-Int Ed 48(23):4194–4197. doi:10.1002/anie.200901399
43. Sun J, Ren J, Zhang S, Cheng W (2009) Water as an efficient medium for the synthesis of
cyclic carbonate. Tetrahedron Lett 50(4):423–426. doi:10.1016/j.tetlet.2008.11.034
44. Yang Z-Z, He L-N, Miao C-X, Chanfreau S (2010) Lewis basic ionic liquids-catalyzed
conversion of carbon dioxide to cyclic carbonates. Adv Synth Catal 352(13):2233–2240.
doi:10.1002/adsc.201000239
45. Aoyagi N, Furusho Y, Endo T (2012) Remarkably efficient catalysts of amidine hydroiodides
for the synthesis of cyclic carbonates from carbon dioxide and epoxides under mild conditions.
Chem Lett 41(3):240–241. doi:10.1246/cl.2012.240
46. Aoyagi N, Furusho Y, Endo T (2013) Convenient synthesis of cyclic carbonates from CO2
and epoxides by simple secondary and primary ammonium iodides as metal-free catalysts
under mild conditions and its application to synthesis of polymer bearing cyclic carbonate
moiety. J Polymer Sci Part A-Polymer Chem 51(5):1230–1242. doi:10.1002/pola.26492
47. Sun J, Han L, Cheng W, Wang J, Zhang X, Zhang S (2011) Efficient acid-base bifunctional
catalysts for the fixation of CO2 with epoxides under metal- and solvent-free conditions.
Chemsuschem 4(4):502–507. doi:10.1002/cssc.201000305
32 M.H. Anthofer et al.

48. Han L, Choi S-J, Park M-S, Lee S-M, Kim Y-J, Kim M-I, Liu B, Park D-W (2012) Carboxylic
acid functionalized imidazolium-based ionic liquids: efficient catalysts for cycloaddition of
CO2 and epoxides. React Kinet Mech Catal 106(1):25–35. doi:10.1007/s11144-011-0399-8
49. Foltran S, Alsarraf J, Robert F, Landais Y, Cloutet E, Cramail H, Tassaing T (2013) On the
chemical fixation of supercritical carbon dioxide with epoxides catalyzed by ionic salts: an in
situ FTIR and Raman study. Catal Sci Technol 3(4):1046–1055. doi:10.1039/c2cy20784b
50. Xiao L, Lv D, Wu W (2011) Brønsted acidic ionic liquids mediated metallic salts catalytic
system for the chemical fixation of carbon dioxide to form cyclic carbonates. Catal Lett 141
(12):1838–1844. doi:10.1007/s10562-011-0682-3
51. Whiteoak CJ, Nova A, Maseras F, Kleij AW (2012) Merging sustainability with
organocatalysis in the formation of organic carbonates by using CO2 as a feedstock.
ChemSusChem 5(10):2032–2038. doi:10.1002/cssc.201200255
52. Wu Z, Xie H, Yu X, Liu E (2013) Lignin-based green catalyst for the chemical fixation of
carbon dioxide with epoxides to form cyclic carbonates under solvent-free conditions.
ChemCatChem. doi:10.1002/cctc.201200894
53. Zhao Y, Yao C, Chen G, Yuan Q (2013) Highly efficient synthesis of cyclic carbonate with CO2
catalyzed by ionic liquid in a microreactor. Green Chem 15(2):446–452. doi:10.1039/c2gc36612f
54. Yu T, Weiss RG (2012) Syntheses of cyclic carbonates with amidinium halide catalysts in
reusable, reversible, room-temperature ionic liquids or acetonitrile. Green Chem 14
(1):209–216. doi:10.1039/c1gc16027c
55. Chatelet B, Joucla L, Dutasta J-P, Martinez A, Szeto KC, Dufaud V (2013) Azaphosphatranes
as structurally tunable organocatalysts for carbonate synthesis from CO2 and epoxides. J Am
Chem Soc 135(14):5348–5351. doi:10.1021/ja402053d
56. Wang J-Q, Kong D-L, Chen J-Y, Cai F, He L-N (2006) Synthesis of cyclic carbonates from
epoxides and carbon dioxide over silica-supported quaternary ammonium salts under super-
critical conditions. J Mol Catal A Chem 249(1–2):143–148, http://dx.doi.org/10.1016/j.
molcata.2006.01.008
57. Wang J-Q, Yue X-D, Cai F, He L-N (2007) Solventless synthesis of cyclic carbonates from
carbon dioxide and epoxides catalyzed by silica-supported ionic liquids under supercritical
conditions. Catal Commun 8(2):167–172, http://dx.doi.org/10.1016/j.catcom.2006.05.049
58. Zhu A, Jiang T, Han B, Zhang J, Xie Y, Ma X (2007) Supported choline chloride/urea as a
heterogeneous catalyst for chemical fixation of carbon dioxide to cyclic carbonates. Green
Chem 9(2):169–172. doi:10.1039/B612164K
59. Xie H, Duan H, Li S, Zhang S (2005) The effective synthesis of propylene carbonate catalyzed by
silica-supported hexaalkylguanidinium chloride. New J Chem 29(9):1199–1203. doi:10.1039/
B504822B
60. Lai G, Peng J, Li J, Qiu H, Jiang J, Jiang K, Shen Y (2006) Ionic liquid functionalized silica
gel: novel catalyst and fixed solvent. Tetrahedron Lett 47(39):6951–6953, http://dx.doi.org/
10.1016/j.tetlet.2006.07.122
61. Zhang X, Wang D, Zhao N, Al-Arifi ASN, Aouak T, Al-Othman ZA, Wei W, Sun Y (2009)
Grafted ionic liquid: catalyst for solventless cycloaddition of carbon dioxide and propylene
oxide. Catal Commun 11(1):43–46, http://dx.doi.org/10.1016/j.catcom.2009.08.007
62. Sakai T, Tsutsumi Y, Ema T (2008) Highly active and robust organic-inorganic hybrid
catalyst for the synthesis of cyclic carbonates from carbon dioxide and epoxides. Green
Chem 10(3):337–341. doi:10.1039/B718321F
63. Takahashi T, Watahiki T, Kitazume S, Yasuda H, Sakakura T (2006) Synergistic hybrid
catalyst for cyclic carbonate synthesis: remarkable acceleration caused by immobilization of
homogeneous catalyst on silica. Chem Commun 15:1664–1666. doi:10.1039/B517140G
64. Udayakumar S, Park S-W, Park D-W, Choi B-S (2008) Immobilization of ionic liquid on
hybrid MCM-41 system for the chemical fixation of carbon dioxide on cyclic carbonate. Catal
Commun 9(7):1563–1570, http://dx.doi.org/10.1016/j.catcom.2008.01.001
65. Udayakumar S, Raman V, Shim H-L, Park D-W (2009) Cycloaddition of carbon dioxide for
commercially-imperative cyclic carbonates using ionic liquid-functionalized porous amorphous
silica. Appl Catal A Gen 368(1–2):97–104, http://dx.doi.org/10.1016/j.apcata.2009.08.015
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 33

66. Udayakumar S, Lee M-K, Shim H-L, Park S-W, Park D-W (2009) Imidazolium derivatives
functionalized MCM-41 for catalytic conversion of carbon dioxide to cyclic carbonate. Catal
Commun 10(5):659–664, http://dx.doi.org/10.1016/j.catcom.2008.11.017
67. Dharman MM, Choi H-J, Kim D-W, Park D-W (2011) Synthesis of cyclic carbonate through
microwave irradiation using silica-supported ionic liquids: effect of variation in the silica
support. Catal Today 164(1):544–547, http://dx.doi.org/10.1016/j.cattod.2010.11.009
68. Han L, Park S-W, Park D-W (2009) Silica grafted imidazolium-based ionic liquids: efficient
heterogeneous catalysts for chemical fixation of CO2 to a cyclic carbonate. Energy Environ
Sci 2(12):1286–1292. doi:10.1039/B910763K
69. Han L, Choi H-J, Choi S-J, Liu B, Park D-W (2011) Ionic liquids containing carboxyl acid
moieties grafted onto silica: synthesis and application as heterogeneous catalysts for cyclo-
addition reactions of epoxide and carbon dioxide. Green Chem 13(4):1023–1028. doi:10.
1039/C0GC00612B
70. Appaturi JN, Adam F (2013) A facile and efficient synthesis of styrene carbonate via
cycloaddition of CO2 to styrene oxide over ordered mesoporous MCM-41-Imi/Br catalyst.
Appl Catal B Environ 136–137:150–159, http://dx.doi.org/10.1016/j.apcatb.2013.01.049
71. Motokura K, Itagaki S, Iwasawa Y, Miyaji A, Baba T (2009) Silica-supported aminopyridinium
halides for catalytic transformations of epoxides to cyclic carbonates under atmospheric pressure
of carbon dioxide. Green Chem 11(11):1876–1880. doi:10.1039/B916764C
72. Udayakumar S, Son Y-S, Lee M-K, Park S-W, Park D-W (2008) The synthesis of
chloropropylated MCM-41 through co-condensation technique: the path finding process.
Appl Catal A Gen 347(2):192–199, http://dx.doi.org/10.1016/j.apcata.2008.06.009
73. Cheng W, Chen X, Sun J, Wang J, Zhang S (2013) SBA-15 supported triazolium-based ionic
liquids as highly efficient and recyclable catalysts for fixation of CO2 with epoxides. Catal
Today 200:117–124, http://dx.doi.org/10.1016/j.cattod.2012.10.001
74. Xie Y, Zhang Z, Jiang T, He J, Han B, Wu T, Ding K (2007) CO2 cycloaddition reactions
catalyzed by an ionic liquid grafted onto a highly cross-linked polymer matrix. Angew Chem
Int Ed 46(38):7255–7258. doi:10.1002/anie.200701467
75. Shi T-Y, Wang J-Q, Sun J, Wang M-H, Cheng W-G, Zhang S-J (2013) Efficient fixation of
CO2 into cyclic carbonates catalyzed by hydroxyl-functionalized poly(ionic liquids). RSC
Adv 3(11):3726–3732. doi:10.1039/C3RA21872D
76. Xiong Y, Wang H, Wang R, Yan Y, Zheng B, Wang Y (2010) A facile one-step synthesis to
cross-linked polymeric nanoparticles as highly active and selective catalysts for cycloaddi-
tion of CO2 to epoxides. Chem Commun 46(19):3399–3401. doi:10.1039/B926901K
77. Xiong Y, Wang Y, Wang H, Wang R (2011) A facile one-step synthesis to ionic liquid-based
cross-linked polymeric nanoparticles and their application for CO2 fixation. Polym Chem
2(10):2306–2315. doi:10.1039/C1PY00201E
78. Xiong Y, Wang Y, Wang H, Wang R, Cui Z (2012) Novel one-step synthesis to cross-linked
polymeric nanoparticles as highly active and selective catalysts for cycloaddition of CO2 to
epoxides. J Appl Polym Sci 123(3):1486–1493. doi:10.1002/app.34622
79. Nishikubo T, Kameyama A, Yamashita J, Tomoi M, Fukuda W (1993) Insoluble polystyrene-
bound quaternary onium salt catalysts for the synthesis of cyclic carbonates by the reaction of
oxiranes with carbon dioxide. J Polym Sci A Polym Chem 31(4):939–947. doi:10.1002/pola.
1993.080310412
80. Park D-W, Yu B-S, Jeong E-S, Kim I, Kim M-I, Oh K-J, Park S-W (2004) Comparative
studies on the performance of immobilized quaternary ammonium salt catalysts for the
addition of carbon dioxide to glycidyl methacrylate. Catal Today 98(4):499–504, http://dx.
doi.org/10.1016/j.cattod.2004.09.003
81. Du Y, Cai F, Kong D-L, He L-N (2005) Organic solvent-free process for the synthesis of
propylene carbonate from supercritical carbon dioxide and propylene oxide catalyzed by
insoluble ion exchange resins. Green Chem 7(7):518–523. doi:10.1039/B500074B
82. Ma J, Song J, Liu H, Liu J, Zhang Z, Jiang T, Fan H, Han B (2012) One-pot conversion of
CO2 and glycerol to value-added products using propylene oxide as the coupling agent. Green
Chem 14(6):1743–1748. doi:10.1039/C2GC35150A
34 M.H. Anthofer et al.

83. Chen X, Sun J, Wang J, Cheng W (2012) Polystyrene-bound diethanolamine based ionic
liquids for chemical fixation of CO2. Tetrahedron Lett 53(22):2684–2688, http://dx.doi.org/
10.1016/j.tetlet.2012.03.058
84. Xiong Y, Bai F, Cui Z, Guo N, Wang R (2013) Cycloaddition reaction of carbon dioxide to
epoxides catalyzed by polymer-supported quaternary phosphonium salts. J Chem 2013:9.
doi:10.1155/2013/261378
85. Sun J, Cheng W, Fan W, Wang Y, Meng Z, Zhang S (2009) Reusable and efficient polymer-
supported task-specific ionic liquid catalyst for cycloaddition of epoxide with CO2. Catal
Today 148(3–4):361–367, http://dx.doi.org/10.1016/j.cattod.2009.07.070
86. Yu J-I, Choi H-J, Selvaraj M, Park D-W (2011) Catalytic performance of polymer-supported
ionic liquids in the cycloaddition of carbon dioxide to allyl glycidyl ether. React Kinet Mech
Catal 102(2):353–365. doi:10.1007/s11144-010-0280-1
87. Watile RA, Deshmukh KM, Dhake KP, Bhanage BM (2012) Efficient synthesis of cyclic carbonate
from carbon dioxide using polymer anchored diol functionalized ionic liquids as a highly active
heterogeneous catalyst. Catal Sci Technol 2(5):1051–1055. doi:10.1039/C2CY00458E
88. Dai W-L, Chen L, Yin S-F, Li W-H, Zhang Y-Y, Luo S-L, Au C-T (2010) High-efficiency
synthesis of cyclic carbonates from epoxides and CO2 over hydroxyl ionic liquid catalyst
grafted onto cross-linked polymer. Catal Lett 137(1–2):74–80. doi:10.1007/s10562-010-0346-8
89. Zhang Y, Yin S, Luo S, Au CT (2012) Cycloaddition of CO2 to epoxides catalyzed by
carboxyl-functionalized imidazolium-based ionic liquid grafted onto cross-linked polymer.
Indus Eng Chem Res 51(10):3951–3957. doi:10.1021/ie203001u
90. Ochiai B, Endo T (2007) Polymer-supported pyridinium catalysts for synthesis of cyclic
carbonate by reaction of carbon dioxide and oxirane. J Polym Sci A Polym Chem 45
(23):5673–5678. doi:10.1002/pola.22316
91. He J, Wu T, Zhang Z, Ding K, Han B, Xie Y, Jiang T, Liu Z (2007) Cycloaddition of CO2 to
epoxides catalyzed by polyaniline salts. Chem Eur J 13(24):6992–6997. doi:10.1002/chem.
200700210
92. Han L, Choi H-J, Kim D-K, Park S-W, Liu B, Park D-W (2011) Porous polymer bead-
supported ionic liquids for the synthesis of cyclic carbonate from CO2 and epoxide. J Mol
Catal A Chem 338(1–2):58–64, http://dx.doi.org/10.1016/j.molcata.2011.02.001
93. Cho HC, Lee HS, Chun J, Lee SM, Kim HJ, Son SU (2011) Tubular microporous organic
networks bearing imidazolium salts and their catalytic CO2 conversion to cyclic carbonates.
Chem Commun 47(3):917–919. doi:10.1039/C0CC03914D
94. Song Q-W, He L-N, Wang J-Q, Yasuda H, Sakakura T (2013) Catalytic fixation of CO2 to
cyclic carbonates by phosphonium chlorides immobilized on fluorous polymer. Green Chem
15(1):110–115. doi:10.1039/C2GC36210D
95. Aprile C, Giacalone F, Agrigento P, Liotta LF, Martens JA, Pescarmona PP, Gruttadauria M
(2011) Multilayered supported ionic liquids as catalysts for chemical fixation of carbon dioxide:
a high-throughput study in supercritical conditions. ChemSusChem 4(12):1830–1837.
doi:10.1002/cssc.201100446
96. Du Y, Wang J-Q, Chen J-Y, Cai F, Tian J-S, Kong D-L, He L-N (2006) A poly(ethylene
glycol)-supported quaternary ammonium salt for highly efficient and environmentally
friendly chemical fixation of CO2 with epoxides under supercritical conditions. Tetrahedron
Lett 47(8):1271–1275, http://dx.doi.org/10.1016/j.tetlet.2005.12.077
97. Dou X-Y, Wang J-Q, Du Y, Wang E, He L-N (2007) Guanidinium salt functionalized PEG:
an effective and recyclable homo-geneous catalyst for the synthesis of cyclic carbonates
from CO2 and epoxides under solvent-free conditions. Synlett 2007:3058–3062. doi:10.1055/
s-2007-992362
98. Yang Z-Z, Zhao Y-N, He L-N, Gao J, Yin Z-S (2012) Highly efficient conversion of carbon
dioxide catalyzed by polyethylene glycol-functionalized basic ionic liquids. Green Chem
14(2):519–527. doi:10.1039/C2GC16039K
99. Zhao Y-N, Yang Z-Z, Luo S-H, He L-N (2013) Design of task-specific ionic liquids for
catalytic conversion of CO2 with aziridines under mild conditions. Catal Today 200:2–8,
http://dx.doi.org/10.1016/j.cattod.2012.04.006
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 35

100. Patil YP, Tambade PJ, Jagtap SR, Bhanage BM (2008) Synthesis of 2-oxazolidinones/
2-imidazolidinones from CO2, different epoxides and amino alcohols/alkylene diamines
using BrPh3+P-PEG600-P+Ph3Br as homogenous recyclable catalyst. J Mol Catal A
Chem 289(1–2):14–21, http://dx.doi.org/10.1016/j.molcata.2008.03.019
101. Choi H-J, Selvaraj M, Park D-W (2013) Catalytic performance of immobilized ionic liquid
onto PEG for the cycloaddition of carbon dioxide to allyl glycidyl ether. Chem Eng Sci
100:242–248. doi:http://dx.doi.org/10.1016/j.ces.2012.11.014
102. Zhao Y, Tian J-S, Qi X-H, Han Z-N, Zhuang Y-Y, He L-N (2007) Quaternary ammonium
salt-functionalized chitosan: an easily recyclable catalyst for efficient synthesis of cyclic
carbonates from epoxides and carbon dioxide. J Mol Catal A Chem 271(1–2):284–289,
http://dx.doi.org/10.1016/j.molcata.2007.03.047
103. Tharun J, Hwang Y, Roshan R, Ahn S, Kathalikkattil AC, Park D-W (2012) A novel approach
of utilizing quaternized chitosan as a catalyst for the eco-friendly cycloaddition of epoxides
with CO2. Catal Sci Technol 2(8):1674–1680. doi:10.1039/C2CY20137B
104. Tharun J, Kim DW, Roshan R, Hwang Y, Park D-W (2013) Microwave assisted preparation
of quaternized chitosan catalyst for the cycloaddition of CO2 and epoxides. Catal Commun
31:62–65, http://dx.doi.org/10.1016/j.catcom.2012.11.018
105. Sun J, Wang J, Cheng W, Zhang J, Li X, Zhang S, She Y (2012) Chitosan functionalized ionic
liquid as a recyclable biopolymer-supported catalyst for cycloaddition of CO2. Green Chem
14(3):654–660. doi:10.1039/C2GC16335G
106. Roshan KR, Mathai G, Kim J, Tharun J, Park G-A, Park D-W (2012) A biopolymer mediated
efficient synthesis of cyclic carbonates from epoxides and carbon dioxide. Green Chem
14(10):2933–2940. doi:10.1039/C2GC35942A
107. Barbarini A, Maggi R, Mazzacani A, Mori G, Sartori G, Sartorio R (2003) Cycloaddition of
CO2 to epoxides over both homogeneous and silica-supported guanidine catalysts. Tetrahe-
dron Lett 44(14):2931–2934, http://dx.doi.org/10.1016/S0040-4039(03)00424-6
108. Zhang X, Zhao N, Wei W, Sun Y (2006) Chemical fixation of carbon dioxide to propylene
carbonate over amine-functionalized silica catalysts. Catal Today 115(1–4):102–106,
http://dx.doi.org/10.1016/j.cattod.2006.02.028
109. Zhang X, Zhang Y, Yang Y, Wei Q, Zhang X (2008) Chemical fixation of carbon dioxide
to propylene carbonate over TBD/SiO2 and DBU/SiO2 catalysts. React Kinet Catal Lett
94(2):385–390. doi:10.1007/s11144-008-5275-9
110. Yu KMK, Curcic I, Gabriel J, Morganstewart H, Tsang SC (2009) Catalytic coupling of CO2
with epoxide over supported and unsupported amines. J Phys Chem A 114(11):3863–3872.
doi:10.1021/jp906365g
111. Sankar M, Tarte NH, Manikandan P (2004) Effective catalytic system of zinc-substituted
polyoxometalate for cycloaddition of CO2 to epoxides. Appl Cataly A Gen 276(1–2):217–222,
http://dx.doi.org/10.1016/j.apcata.2004.08.008
112. Shiels RA, Jones CW (2007) Homogeneous and heterogeneous 4-(N, N-dialkylamino)pyri-
dines as effective single component catalysts in the synthesis of propylene carbonate. J Mol
Catal A Chem 261(2):160–166, http://dx.doi.org/10.1016/j.molcata.2006.08.002
113. Srivastava R, Srinivas D, Ratnasamy P (2005) CO2 activation and synthesis of cyclic
carbonates and alkyl/aryl carbamates over adenine-modified Ti-SBA-15 solid catalysts.
J Catal 233(1):1–15, http://dx.doi.org/10.1016/j.jcat.2005.03.023
114. Srivastava R, Srinivas D, Ratnasamy P (2006) Sites for CO2 activation over amine-functionalized
mesoporous Ti(Al)-SBA-15 catalysts. Micropor Mesopor Mater 90(1–3):314–326, http://dx.doi.
org/10.1016/j.micromeso.2005.10.043
115. Jagtap SR, Raje VP, Samant SD, Bhanage BM (2007) Silica supported polyvinyl pyridine as
a highly active heterogeneous base catalyst for the synthesis of cyclic carbonates from carbon
dioxide and epoxides. J Mol Catal A Chem 266(1–2):69–74, http://dx.doi.org/10.1016/j.
molcata.2006.10.033
116. Adam F, Batagarawa MS (2013) Tetramethylguanidine–silica nanoparticles as an efficient and
reusable catalyst for the synthesis of cyclic propylene carbonate from carbon dioxide and
propylene oxide. Appl Catal A Gen 454:164–171, http://dx.doi.org/10.1016/j.apcata.2012.12.009
36 M.H. Anthofer et al.

117. Prasetyanto EA, Ansari MB, Min B-H, Park S-E (2010) Melamine tri-silsesquioxane bridged
periodic mesoporous organosilica as an efficient metal-free catalyst for CO2 activation. Catal
Today 158(3–4):252–257, http://dx.doi.org/10.1016/j.cattod.2010.03.081
118. Roeser J, Kailasam K, Thomas A (2012) Covalent triazine frameworks as heterogeneous
catalysts for the synthesis of cyclic and linear carbonates from carbon dioxide and epoxides.
ChemSusChem 5(9):1793–1799. doi:10.1002/cssc.201200091
119. Huang J-W, Shi M (2003) Chemical fixation of carbon dioxide by NaI/PPh3/PhOH. J Org
Chem 68(17):6705–6709. doi:10.1021/jo0348221
120. Shen Y-M, Duan W-L, Shi M (2003) Phenol and organic bases co-catalyzed chemical
fixation of carbon dioxide with terminal epoxides to form cyclic carbonates. Adv Synth
Catal 345(3):337–340. doi:10.1002/adsc.200390035
121. Wang J-Q, Sun J, Cheng W-G, Dong K, Zhang X-P, Zhang S-J (2012) Experimental and
theoretical studies on hydrogen bond-promoted fixation of carbon dioxide and epoxides in
cyclic carbonates. Phys Chem Chem Phys 14(31):11021–11026. doi:10.1039/c2cp41698k
122. Liang S, Liu H, Jiang T, Song J, Yang G, Han B (2011) Highly efficient synthesis of cyclic
carbonates from CO2 and epoxides over cellulose/KI. Chem Commun 47(7):2131–2133.
doi:10.1039/c0cc04829a
123. Song J, Zhang Z, Han B, Hu S, Li W, Xie Y (2008) Synthesis of cyclic carbonates from
epoxides and CO2 catalyzed by potassium halide in the presence of [small beta]-cyclodextrin.
Green Chem 10(12):1337–1341. doi:10.1039/B815105A
124. Song J, Zhang B, Zhang P, Ma J, Liu J, Fan H, Jiang T, Han B (2012) Highly efficient
synthesis of cyclic carbonates from CO2 and epoxides catalyzed by KI/lecithin. Catal Today
183(1):130–135. doi:10.1016/j.cattod.2011.08.042
125. Zhou L, Liu Y, He Z, Luo Y, Zhou F, Yu E, Hou Z, Eli W (2013) Pentaerythritol and KI:
an efficient catalytic system for the conversion from CO2 and epoxides to cyclic carbonates.
J Chem Res 2:102–104. doi:10.3184/174751913x13571500195988
126. Ma J, Liu J, Zhang Z, Han B (2012) The catalytic mechanism of KI and the co-catalytic
mechanism of hydroxyl substances for cycloaddition of CO2 with propylene oxide. Green
Chem 14(9):2410–2420. doi:10.1039/c2gc35711a
127. Sonnati MO, Amigoni S, de Givenchy EPT, Darmanin T, Choulet O, Guittard F (2013)
Glycerol carbonate as a versatile building block for tomorrow: synthesis, reactivity, properties
and applications. Green Chem 15(2):283–306. doi:10.1039/c2gc36525a
128. Cui H, Wang T, Wang F, Gu C, Wang P, Dai Y (2003) One-pot synthesis of dimethyl
carbonate using ethylene oxide, methanol, and carbon dioxide under supercritical conditions.
Indus Eng Chem Res 42(17):3865–3870. doi:10.1021/ie021014b
129. Cui H, Wang T, Wang F, Gu C, Wang P, Dai Y (2004) Kinetic study on the one-pot synthesis of
dimethyl carbonate in supercritical CO2 conditions. Indus Eng Chem Res 43(24):7732–7739.
doi:10.1021/ie049715r
130. Kishimoto Y, Ogawa I (2004) Amine-catalyzed, one-pot coproduction of dialkyl carbonates and
1,2-diols from epoxides, alcohols, and carbon dioxide. Indus Eng Chem Res 43(26):8155–8162.
doi:10.1021/ie040006n
131. Chen X, Hu C, Su J, Yu T, Gao Z (2006) One-pot synthesis of dimethyl carbonate catalyzed
by [bmim]BF4/CH3ONa. Chin J Catal 27(6):485–488, http://dx.doi.org/10.1016/S1872-2067
(06)60029-6
132. Tian J-S, Wang J-Q, Chen J-Y, Fan J-G, Cai F, He L-N (2006) One-pot synthesis of dimethyl
carbonate catalyzed by n-Bu4NBr/n-Bu3N from methanol, epoxides, and supercritical CO2.
Appl Catal A Gen 301(2):215–221, http://dx.doi.org/10.1016/j.apcata.2005.12.002
133. Tian J-S, Miao C-X, Wang J-Q, Cai F, Du Y, Zhao Y, He L-N (2007) Efficient synthesis of
dimethyl carbonate from methanol, propylene oxide and CO2 catalyzed by recyclable inorganic
base/phosphonium halide-functionalized polyethylene glycol. Green Chem 9(6):566–571.
doi:10.1039/B614259A
134. Li J, Wang L, Shi F, Liu S, He Y, Lu L, Ma X, Deng Y (2011) Quaternary ammonium ionic
liquids as bi-functional catalysts for one-step synthesis of dimethyl carbonate from ethylene
oxide, carbon dioxide and methanol. Catal Lett 141(2):339–346. doi:10.1007/s10562-010-0498-6
1 Valorization of Carbon Dioxide to Organic Products with Organocatalysts 37

135. Wang J-Q, Sun J, Shi C-Y, Cheng W-G, Zhang X-P, Zhang S-J (2011) Synthesis of dimethyl
carbonate from CO2 and ethylene oxide catalyzed by K2CO3-based binary salts in the
presence of H2O. Green Chem 13(11):3213–3217. doi:10.1039/C1GC15812K
136. Fang S, Fujimoto K (1996) Direct synthesis of dimethyl carbonate from carbon dioxide and
methanol catalyzed by base. Appl Catal Gen 142(1):1–3. doi:10.1016/0926-860X(96)00081-6
137. S-i F, Bhanage BM, Ikushima Y, Arai M (2001) Synthesis of dimethyl carbonate from carbon
dioxide and methanol in the presence of methyl iodide and base catalysts under mild
conditions: effect of reaction conditions and reaction mechanism. Green Chem 3(2):87–91.
doi:10.1039/B100363L
138. Cai Q, Jin C, Lu B, Tangbo H, Shan Y (2005) Synthesis of dimethyl carbonate from methanol
and carbon dioxide using potassium methoxide as catalyst under mild conditions. Catal Lett
103(3–4):225–228. doi:10.1007/s10562-005-7158-2
139. Wang H, Lu B, Cai QH, Wu F, Shan YK (2005) Synthesis of dimethyl carbonate from
methanol and carbon dioxide catalyzed by potassium hydroxide under mild conditions. Chin
Chem Lett 16(9):1267
140. Cai Q, Lu B, Guo L, Shan Y (2009) Studies on synthesis of dimethyl carbonate from
methanol and carbon dioxide. Catal Commun 10(5):605–609, http://dx.doi.org/10.1016/j.
catcom.2008.11.002
141. Riduan SN, Zhang Y, Ying JY (2009) Conversion of carbon dioxide into methanol with
silanes over N-heterocyclic carbene catalysts. Angew Chem Int Ed 48(18):3322–3325.
doi:10.1002/anie.200806058
142. Huang F, Lu G, Zhao L, Li H, Wang Z-X (2010) The catalytic role of N-heterocyclic carbene
in a metal-free conversion of carbon dioxide into methanol: a computational mechanism
study. J Am Chem Soc 132(35):12388–12396. doi:10.1021/ja103531z
143. Das Neves Gomes C, Jacquet O, Villiers C, Thuéry P, Ephritikhine M, Cantat T (2012)
A diagonal approach to chemical recycling of carbon dioxide: organocatalytic transformation
for the reductive functionalization of CO2. Angew Chem 124(1):191–194. doi:10.1002/ange.
201105516
144. Jacquet O, Das Neves Gomes C, Ephritikhine M, Cantat T (2012) Recycling of carbon and
silicon wastes: room temperature formylation of N–H bonds using carbon dioxide and
polymethylhydrosiloxane. J Am Chem Soc 134(6):2934–2937. doi:10.1021/ja211527q
145. Jacquet O, Das Neves Gomes C, Ephritikhine M, Cantat T (2013) Complete catalytic
deoxygenation of CO2 into formamidine derivatives. ChemCatChem 5(1):117–120. doi:10.
1002/cctc.201200732
146. Stephan DW, Erker G (2010) Frustrated Lewis pairs: metal-free hydrogen activation and
more. Angew Chem Int Ed 49(1):46–76. doi:10.1002/anie.200903708
147. Ashley AE, Thompson AL, O’Hare D (2009) Non-metal-mediated homogeneous hydrogenation
of CO2 to CH3OH. Angew Chem Int Ed 48(52):9839–9843. doi:10.1002/anie.200905466
148. Berkefeld A, Piers WE, Parvez M (2010) Tandem frustrated Lewis pair/tris(pentafluorophenyl)
borane-catalyzed deoxygenative hydrosilylation of carbon dioxide. J Am Chem Soc
132(31):10660–10661. doi:10.1021/ja105320c
Chapter 2
Direct Transformation of Carbon Dioxide
to Value-Added Products over
Heterogeneous Catalysts

Shin-ichiro Fujita, Masahiko Arai, and Bhalchandra M. Bhanage

2.1 Introduction

Carbon dioxide (CO2) is one of the greenhouse gases arising from human activities,
and hence its accumulation in the atmosphere should be controlled by removing it
from industrial emissions to avoid global warming. Thus, large amounts of CO2
would be available [1–4]. On the other hand, CO2 is recognized to be an abundant,
cheap, recyclable, and nontoxic carbon source that can sometimes replace toxic
chemicals such as phosgene, isocyanates, or carbon monoxide [2, 5–9]. It has been
considered that utilization of CO2 is more attractive rather than storage, if economical
processes are available. Under these circumstances, chemical fixation of CO2 into
valuable chemicals is still gaining considerable interest.
Heterogeneous catalysts are widely used in industries because of their advantages
in the separation and recycling and their applicability for flow reaction systems.
Unfortunately, conventional heterogeneous catalysts are sometimes less active and/or
selective in comparison to such homogeneous ones as metal complexes and organic
bases including ionic liquids. A class of heterogeneous catalysts can be built of those
homogeneous ones immobilized on solids of metal oxides or polymers. These
immobilized catalysts can have not only high activity and/or selectivity but also the
advantages in the catalyst separation, enabling more efficient and economical reaction
processes.
During the last two decades, a variety of catalysts have been employed for the
chemical fixation of CO2. In this chapter, the utilization of heterogeneous catalysts for

S.-i. Fujita (*) • M. Arai


Division of Chemical Process Engineering, Faculty of Engineering,
Hokkaido University, Sapporo 060-8628, Japan
e-mail: sfuji@eng.hokudai.ac.jp; marai@eng.hokudai.ac
B.M. Bhanage
Department of Chemistry, Institute of Chemical Technology,
Mumbai 400019, India
e-mail: bhanage@ictmumbai.edu.in

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 39
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_2,
© Springer-Verlag Berlin Heidelberg 2014
40 S.-i. Fujita et al.

the synthesis of cyclic carbonate from CO2 and epoxides and of dimethyl carbonate
(DMC) from CO2 and methanol has been reviewed. The catalysts described will include
immobilized organic base catalysts as well as conventional inorganic solid catalysts of
metal oxides, clays, and zeolites. A few related reactions will also be described.

2.2 Synthesis of Cyclic Carbonate via Cycloaddition


of CO2 to Epoxide

2.2.1 Use of Inorganic Catalysts for CO2 Cycloaddition

The five-member cyclic carbonates have several applications [6–8]; e.g., ethylene
carbonate is an excellent solvent for many polymers and resins. Another important
application includes the use as an intermediate for the manufacture of several
important chemicals like dialkyl carbonates, glycols, carbamates, pyrimidines,
purines, etc. These five-member cyclic carbonates were conventionally produced
via a corrosive, poisonous, and hazardous route involving glycol and phosgene. Then
manufacturers shifted to the production based on the cycloaddition of CO2 to epoxide
(Scheme 2.1) in the presence of base catalysts.
Yano et al. first reported the use of a metal oxide catalyst for the reaction of CO2
and epoxide. They showed that a commercial MgO could catalyze the reactions of
CO2 with propylene oxide (PO) and styrene oxide (SO) to give cyclic carbonates
[10]. It gave 60 % and 41 % yields of styrene carbonate (SC) and propylene
carbonate (PC), respectively, at 135  C and 2 MPa of CO2 for 12 h in a solvent
of dimethylformamide (DMF). CO2 pressure showed no effect on the carbonate
yield in a region between 0.2 and 2.8 MPa. When a chiral epoxide was used as the
substrate, interestingly, the chirality was retained in the cyclic carbonate product.
Yamaguchi et al. used Mg–Al mixed oxides prepared from hydrotalcite for the
reactions of CO2 with several epoxides [11]. Corresponding cyclic carbonates were
obtained with excellent yields at 100–120  C and 0.5 MPa of CO2 for 15–24 h in
DMF. The Mg/Al ratio of the mixed oxide was significant for its catalytic activity,
and the best value was 5. Based on the catalyst characterization results, the authors
proposed that the O and Al atoms involved in Mg–O–Al bonds act as basic and
acidic sites, respectively, and they cooperatively activate an epoxide molecule.

Scheme 2.1 Synthesis of O


cyclic carbonate via O
cycloaddition of CO2 to + CO2 O O
epoxide. Abbreviations for
R
a few representative
R
substrates and products are
also given R = H (EO: ethylene oxide, EC: ethylene carbonate)
CH3 (PO: propylene oxide, PC: propylene carbonate)
C6H5 (SO: styrene oxide, SC: styrene carbonate)
2 Direct Transformation of Carbon Dioxide to Value-Added Products. . . 41

No use of organic solvent makes the reaction system more preferable. A few
groups have reported the CO2 cycloaddition reactions under such conditions.
Bhanage et al. employed several metal oxide catalysts of MgO, CaO, ZnO, Zr2O,
La2O3, CeO2, and Al2O3 for the reaction of CO2 and propylene oxide (PO) to
propylene carbonate (PC) [12]. MgO was found to be the best among the metal
oxides examined. It gave a PC yield of 32 % with a PC selectivity of 92 % at 150  C
and 8 MPa of CO2 for 15 h. Although a higher PC yield of 54 % was obtained with
La2O3, the PC selectivity was lower (75 %). Basic properties of these catalysts were
investigated by temperature programmed desorption (TPD) of adsorbed CO2. It was
suggested that both moderately and strongly basic sites are active sites for the
reaction, and MgO has a large amount of moderately basic sites, resulting in its high
activity and selectivity. The reaction of CO2 and styrene oxide (SO) to styrene
carbonate (SC) was also examined with MgO. For this reaction, MgO gave a poor
product yield (16 %) because of low selectivity.
Tu and Davis prepared cesium-loaded zeolite X (Cs/KX) catalysts and used them
for the synthesis of ethylene carbonate (EC) from CO2 and ethylene oxide
(EO) [13]. The activity of the zeolite catalyst was slightly lower than MgO, but the
former catalyst was much more tolerant of water moisture than the latter one. Wet
Cs/KX selectively produced EC, while wet MgO mostly produced mono-, di-, and
triethylene glycol, which were formed by hydrolysis of EO and following dehydra-
tion reactions. Interestingly, the wet Cs/KX afforded a much higher EC yield than dry
Cs/KX, although the reason was unknown.
Smectite is one of the layered clay minerals, and its acidic and basic properties are
tunable [14]. Fujita et al. developed new types of catalysts based on Mg- and/or
Ni-containing smectite catalysts, in which various amounts of alkali metals such as
Na, K, and Li are incorporated [15, 16]. These catalysts were highly active and
selective for PC synthesis from CO2 and PO without organic solvent. The catalytic
activity strongly depends on the amount of alkali atoms incorporated. The most active
smectite catalyst gave a PC yield of 81 % with a PC selectivity of 94 % at 150  C and
8 MPa of CO2 for 15 h. The turnover number obtained with the most active smectite
catalyst was about five times of that obtained with MgO, which was considered to be
the best catalyst among the conventional metal oxide catalysts tested, as stated above.
TPD of adsorbed CO2 suggested that the smectite catalysts had large amounts of
strongly basic sites, resulting in their high activities. They also studied in detail the
optimization of the reaction conditions for this reaction [16]. The CO2 pressure does
not significantly affect the conversion and the selectivity in the region between 3 and
10 MPa; however, both the conversion and the selectivity decrease sharply at 15 MPa.
It is highly probable that the reaction system consists of three phases: CO2–liquid–solid
at the lower pressures, while it should be in a homogeneous phase at 15 MPa. Such a
phase change would cause an increase in the volume at the location where the reaction
proceeds. Hence, the concentration of the catalyst and/or PO would be low at 15 MPa,
resulting in low PO conversion at this elevated pressure. The effect of reaction
temperature on the reaction of PO and CO2 was investigated. The PO conversion
was found to increase with the temperature, as expected; however, the selectivity
decreased at 170  C and thus the optimum reaction temperature was 150  C.
42 S.-i. Fujita et al.

O
A O O
A C
O A
O
O− O O O− O O
+ +
R R
R B R B R
A = Metal cation B
B = Base

Scheme 2.2 A generally accepted reaction mechanism for CO2 cycloaddition to epoxide catalyzed
by acid-base bi-functional systems

As mentioned above, MgO gave a reasonable SC yield from CO2 and SO at


135  C in the presence of DMF; however, it gave a poor SC yield even at 150  C in
the absence of DMF, suggesting positive effects of DMF on the reaction. Yasuda
et al. carried out the PC synthesis with MgO, Mg–Al mixed oxide, or SmOCl in the
presence and absence of DMF and showed that the presence of DMF enhanced the
PC yield twice or more irrespective of the catalyst used [17]. Similar results were
also reported for the cyclic carbonate synthesis using Nb2O3 and DMF and using
titanosilicate and 4-(dimethylamino)pyridine (DMAP) [18, 19]. Furthermore, Mori
et al. reported that, when even a catalytic amount of DMAP was combined with
zinc-based hydroxyapatite (ZnHAP), a good SC yield of 79 % was obtained at
100  C and 0.5 MPa of CO2 for 20 h, although ZnHAP had no activity and the
activity of DMAP was very low [20]. Thus, it is highly probable that the occurrence
of the synergetic effect is common in the bifunctional catalyst systems consisting of
inorganic metal compounds and organic bases. A general reaction mechanism for
those bifunctional catalyst systems can be drawn as Scheme 2.2. The Lewis acidic
center of metal cation would interact with the oxygen atom of the epoxide, and
the base would attack the less hindered carbon atom of the epoxide ring. Such
cooperative activation of the epoxide should make the ring opening easier, being
the reason for the promotional effects of the base. It is also possible that base
activates a CO2 molecule. In the absence of organic bases, the oxygen atoms on the
metal oxide surface would act as Lewis base, but it would be less effective than the
organic base.

2.2.2 Use of Immobilized Organic Base Catalysts


for CO2 Cycloaddition

It is well known that organic bases of amines, ammonium salts, phosphonium salts,
imidazolium salts, etc., can be homogeneous catalysts for the cyclic carbonate
synthesis via CO2 cycloaddition to epoxides [7, 8, 21, 22]. Several groups examined
the immobilization of those catalysts onto a solid support to facilitate the catalyst
separation from the products. In this subsection, some representative examples of
the CO2 cycloaddition reactions using such immobilized catalysts are described.
2 Direct Transformation of Carbon Dioxide to Value-Added Products. . . 43

NH 2 NH2
C H 2) 3
) Si( NH2
(EtO 3
SBA-15 N
N N NH
Ti-SBA-15
(EtO N
)3 Si( N
CH ) H
2 3 Cl Cl N N
H
N

Scheme 2.3 Immobilization of propylamine and adenine on SBA-15 or Ti-SBA-15

To synthesize zeolites, tetraalkyl ammonium salts are used as the template.


Srivastava et al. reported that as-synthesized zeolite-beta, which contained the
template of tetramethylammonium bromide, could produce a few types of cyclic
carbonates via CO2 cycloaddition with very high yields at 120  C and 0.7 MPa of
CO2 for 3–8 h without any solvent [23]. The catalyst could be recycled eight times
with little loss in activity. The turnover frequency (TOF) value based on the amount of
the ammonium template contained was almost the same with that of homogeneously
used ammonium bromide. However, the authors pointed out that over the
as-synthesized zeolite-beta, the reaction takes place only at the surface, and therefore,
the ammonium molecules in the inside part of the zeolite are not accessible for the
reaction. Thus, it is highly probable that the actual TOF of the solid catalyst is much
higher than the observed one. This may be explained by the presence of surface silanol
group as described below.
The same group also reported the covalent immobilization of organic amines
(propylamine, adenine) on mesoporous materials of SBA-15 and Ti-SBA-15 by
using silane coupling reagents (Scheme 2.3) [24]. The catalytic activity of the
immobilized catalyst for CO2 cycloaddition increased with the Ti content. FTIR
spectra of the catalysts measured under several conditions suggested the activation
of CO2 on the base sites, because the cyclic carbonate yield was well correlated
with the peak intensity of surface carbamate species. The authors proposed that the
coexistence of epoxide activated on Ti sites and CO2 activated on base sites is
significant for the reaction. Recycling of adenine-modified Ti-SBA-15 was tested.
In the absence of any organic solvent, the activity gradually decreased with
recycling, while such deactivation was much less in the presence of acetonitrile
solvent. When the deactivated catalyst was washed with organic solvents, the initial
activity was almost restored. It was also confirmed that some heavy carbonaceous
products accumulated on the catalyst surface, resulting in the deactivation. So, at
least for this catalyst, the use of organic solvent is required.
The advantage of immobilizing homogeneous catalysts on solid materials is not
only the easier separation of the catalyst but also the usability of it for the
continuous operation using a fixed-bed flow reactor. Takahashi et al. first reported
the use of silica-immobilized tetraalkyl phosphonium bromide for a flow reactor
[25]. The phosphonium bromide was immobilized on silica in a similar way to
Scheme 2.3. They carried out the PC synthesis using 10 MPa of CO2 for more than
1,000 h. Unfortunately, the reaction temperature was required to be increased from
the initial one of 90–160  C for keeping the yield above 80 %, suggesting that some
44 S.-i. Fujita et al.

Scheme 2.4 Cooperative


activation of an epoxide
molecule on silica-
R R
immobilized phosphonium X− O
halides P+
R H

Si Si

SiO2

leaching of the phosphonium bromide from the support might occur; however, the
selectivity to PC was kept over 99.9 % during the whole of the reaction run. In the
same paper [25], they reported interesting synergistic effects of the immobilization
of phosphonium halides on silica. They also carried out the reaction runs in a batch
reactor and showed that the TOF of the immobilized phosphonium halides was
more than 300 times larger compared to that of corresponding phosphonium halides
that were used homogeneously. However, such enhancement in the activity was not
observed by the immobilization on polystyrene. The surprising enhancement of the
activity by supporting on silica was ascribed to the cooperational activation of the
epoxide by the acidic surface silanol groups and the halide anion of the immobilized
phosphonium salts (Scheme 2.4). The most active immobilized catalyst could
almost quantitatively convert PO to PC at 100  C and 10 MPa of CO2 for a short
reaction time of 1 h.
The acidity of the OH group on the solid surface was also suggested to affect the
activity of immobilized onium halide catalysts. Sakai et al. prepared immobilized
phosphonium halide catalysts using silica, aluminosilicate, and basic alumina as the
support materials [26]. The activity for the CO2 cycloaddition to epoxyhexane was
in the order of silica > aluminosilicate > basic alumina, which is the same order as
that of the acidity of the OH group on the support surface, and hence they ascribed
the differences in the activity among the catalysts to those in the acidity of the OH
group of the supports. This also supports the cooperative activation mechanism
shown in Scheme 2.4. The enhancement of the activity by the immobilization
was also observed with imidazolium and aminopyridinium halides [27, 28]. Thus,
the synergistic effects of the immobilization on inorganic materials having surface
hydroxyl groups are common. Over the immobilized catalysts, the mean pore
size, the total pore volume, and the amount of surface hydroxyl groups were
also found to be important factors determining the catalyst performance for CO2
cycloaddition [26, 29].
Zinc tetrahalide onium salts prepared from zinc halide and onium halide were
found to be excellent catalysts for the CO2 cycloaddition [30–32]. They could give
high cyclic carbonate yields under very mild conditions. This is ascribed to
the cooperational activation of epoxide ring by Lewis acidic Zn center and basic
halide anion, which is similar to Scheme 2.2. Qiao et al. applied this strategy for
2 Direct Transformation of Carbon Dioxide to Value-Added Products. . . 45

Scheme 2.5 Activation of +


an epoxide molecule by N N O
PS

polystyrene-supported X O H
diol-functionalized H
imidazolium halide O
R

polymer-immobilized imidazolium catalysts [33]. They immobilized imidazolium


on polystyrene by copolymerization of 1-vinyl-3-butylimidazolium salts and
styrene and further treated the copolymers with zinc halides. The resulting
immobilized catalysts could show high TOF values up to about 3,800 h 1 at
120  C. This TOF value was comparable to those obtained with very active
homogeneous catalyst systems reported so far.
Very recently, Watile et al. developed diol-functionalized imidazolium halide
immobilized on polystyrene (PS) as the catalyst for the CO2 cycloaddition and
showed that the diol-functionalization is more effective for the reaction than the mono
hydroxyl-functionalization [34]. The PC yield obtained with 1-(2,3-di-hydroxyl-
propyl)-imidazolium bromide was higher than that with a corresponding mono
hydroxyl-functionalized imidazolium halide of 1-(2-hydroxyl-ethyl)-imidazolium
bromide. It was proposed that the epoxide is activated by chelate-type hydrogen
bonding of its O atom with the vicinal two hydroxyl groups of the functionalized
imidazolium halide (Scheme 2.5).
Recent advance in the preparation of microporous materials is the development
of zeolitic imidazole frameworks (ZIFs), in which metal atoms such as Zn are
linked through N atoms by ditropic imidazolate to form neutral frameworks [35,
36]. ZIFs have uniform micropores, high surface areas, and open porous framework
structures, making them interesting materials for adsorption and catalytic applica-
tions. The co-presence of the Lewis acidic Zn center and the N basic moieties in ZIF
would make ZIF a candidate of the catalyst for the CO2 cycloaddition reaction.
Recently, Miralda et al. prepared non-functionalized and ethylene diamine-
functionalized ZIF-8 (one kind of ZIF) catalysts and used them for the CO2
cycloaddition to epichlorohydrin [37]. The nonfunctionalized ZIF-8 produced
chloropropylene carbonate with a 44 % yield at 80  C and 0.7 MPa of CO2 for
4 h; however, the selectivity to cyclic carbonate was 52 %. By-products were a diol
(24 % selectivity) and a dimer of the substrate (24 % selectivity). At a higher
temperature of 100  C, the conversion was increased, but the carbonate yield was
decreased because of the increase in the dimer formation. The functionalization of
ZIF-8 with amine improved both the yield of and the selectivity to the carbonate
(73 % and 73 %), and the formation of the dimer was ceased. Unfortunately,
the catalytic properties of both nonfunctionalized and functionalized ZIF-8
catalysts were unstable. They lost about half of their initial activities after the
first recycling. This was ascribed to the accumulation of carbonaceous materials
on the catalysts. Furthermore, after the second recycling, the zeolitic structures
of both two catalysts were destroyed. The authors examined ZIF-8 for only the
46 S.-i. Fujita et al.

substrate of epichlorohydrin. It is not clear at present whether the deactivation


occurs with other epoxides. The loss of their structural features was proposed to
result from the co-presence of CO2 at high pressure and the carbonaceous materials.
ZIFs are a new class of porous materials on which the number of studies is limited.
Further studies on them may improve their catalytic performance and stability.

2.3 Synthesis of Dimethyl Carbonate from CO2


and Methanol

Dimethyl carbonate (DMC) finds such applications as a solvent, an octane


enhancer, and a reagent for carbonylation and methylation, replacing poisonous
phosgene and dimethyl sulfate [5, 9, 38]. In addition, DMC is a precursor for
polycarbonate resins, which have been commercially produced by polycondensa-
tion between bisphenol-A and phosgene. Thus, DMC can be regarded as a safe
replacement for phosgene. Currently, DMC is synthesized by phosgenation of
methanol, by oxidative carbonylation of methanol (non-phosgene route), or by a
reaction of carbon monoxide with methyl nitrite. Methyl nitrite is produced from
methanol, oxygen, and nitric oxide. These routes use poisonous and/or corrosive
gases such as phosgene, hydrogen chloride, nitric oxide, and carbon monoxide
and also bear potential explosion hazards in the case of methanol carbonylation.
Hence, the direct synthesis of DMC from CO2 and methanol (Scheme 2.6) is an
attractive alternative and, hence, one of the promising reactions for the chemical
fixation of CO2.
Fang and Fujimoto [39] reported that various inorganic bases catalyze the direct
DMC synthesis in the presence of methyl iodide as a promoter. Among the base
catalysts examined, potassium carbonate was the most effective. Although the yield
of DMC obtained with this catalyst was higher than those with organometallic
complexes, the range of reaction conditions used was quite limited. Fujita
et al. studied the influence of reaction conditions and the reaction mechanism in
detail [40]. It was found that two maxima in the DMC formation appear at CO2
pressures of 4.5 and 8 MPa, while the formation of a by-product, dimethyl ether
(DME), decreased monotonically with increasing CO2 pressure. DMC formation
increased almost linearly with the amount of either methyl iodide or potassium
carbonate when their concentrations were low. Mechanistic studies suggested that
DMC and DME are produced in parallel pathways and methyl iodide is involved in
both the formation of DMC and DME, as illustrated in Scheme 2.7. The catalyst
was deactivated in the course of the reaction and methyl iodide was essentially a
reactant rather than a promoter for the reaction. This is likely because the

O
Scheme 2.6 Synthesis of
DMC from CO2 and 2CH3OH + CO2 + H2O
methanol MeO OMe
2 Direct Transformation of Carbon Dioxide to Value-Added Products. . . 47

CH3OH + Base CH3O− + H+-Base


O
CO2 CH3I
CH3O− [CH3COO]− + I−
MeO OMe
CH3I
MeOMe + I−

H+-Base + I− HI + Base

CH3OH + HI CH3I + H2O

Scheme 2.7 Reaction mechanism of DMC synthesis in the presence of base and CH3I

regeneration of methyl iodide from methanol and HI produced is difficult under


the present reaction conditions. The residual HI instead reacted with potassium
carbonate as confirmed by IR spectroscopy.
The direct DMC synthesis using solid catalysts in the absence of methyl iodide
was also reported. Tomishige et al. found that ZrO2 and CeO2–ZrO2 mixed oxide
catalysts were selective for the direct DMC synthesis reaction under such conditions
[41–43]. CeO2–ZrO2 mixed oxides were more active than pure ZrO2. Based on
temperature programmed desorption of NH3 and CO2, they proposed that acid–base
pairs are active sites for the reaction. Jung and Bell also drew the same conclusion
from IR measurements of the surface species formed on ZrO2 [44, 45]. It was also
reported that the enhancement in the acidity of ZrO2 by supporting phosphoric or
tungstophosphoric acid significantly improved the catalytic activity [46, 47]. Thus,
several ZrO2-based catalysts were disclosed; however, the DMC yields obtained were
far from satisfactory because of the reaction equilibrium limitation. Lately,
Tomishige and Kunimori [48] showed that the DMC yield can be improved by
using dimethyl acetal (DMP) as a dehydration reagent. Hence, removal of H2O
from the reaction system is significant to get DMC in high yields; however, the
amount of DMP added should be controlled. With too much DMP, the DMC
formation decreased and the formation of undesirable by-product dimethyl ether
increased. It was also shown that diethyl carbonate could be synthesized from ethanol
and CO2, although its yield was lower than that of DMC.
An analogous reaction to the DMC synthesis from CO2 and methanol is diphenyl
carbonate (DPC) synthesis from CO2 and phenol. DPC is an important precursor of
aromatic polycarbonate. It is conventionally synthesized from phosgene and phe-
nol. Also for the synthesis of DPC, the replacement of phosgene with CO2 is
significant. Although it has been known that Lewis acid compounds of zinc halides
can catalyze this reaction with or without base in a solvent of CCl4, the use of CCl4
is essential for the reaction, because CCl4 is considered to be transformed to
phosgene in the reaction system [49–51]. One can say that those systems are the
reaction using in situ generated phosgene. Furthermore, the use of toxic CCl4
should be avoided. Li and Su tried the reaction using ZrO2 without any solvent at
an elevated temperature and CO2 pressure [52]. Unfortunately, the catalyst could
48 S.-i. Fujita et al.

produce no DPC. Products observed were hydroxybenzyl phenol, diphenyl ether,


and other compounds. For the DPC synthesis, further development of the catalyst is
required. It may be a key issue how effectively to create a new bond between the
carbon atom of CO2 and the O atom of phenol.

2.4 Other Reactions for Transformation of CO2

In the foregoing sections, the utilization of heterogeneous catalysts for the cyclic
carbonate synthesis from CO2 and epoxide and for the DMC synthesis from CO2
and methanol was reviewed. In this section, a few related reactions in the presence
of heterogeneous catalysts have been described.
Tomishige et al. examined the reaction of CO2 and glycol to produce cyclic
carbonate with CeO2–ZrO2 mixed oxide catalysts (Scheme 2.8) [53]. This is another
route to synthesize cyclic carbonate from CO2. Using the substrates of glycols cheaper
than epoxides is of benefit. They optimized the catalyst composition (Ce/Zr ratio) and
the calcination condition. Even after the optimization, however, the carbonate yields
obtained were very low because of the reaction equilibrium limitation. Removal of
H2O from the reaction system would be essential to obtain high yields of the carbonate,
similar to the cases of direct DMC synthesis from CO2 and methanol.
Carbamates are another important class of organic compounds that meet a
variety of applications for the production of polyurethanes, pesticides, fungicides,
and medicinal drugs. They are commercially produced with the use of toxic
phosgene. Another route producing carbamate is the reaction of primary amine,
CO2, and alkyl halide (Scheme 2.9). It is a benign way to synthesize carbamate
avoiding the use of toxic phosgene, isocyanate, and CO. Srivastava et al. who used
as-synthesized zeolite and amine-modified SBA-15 and Ti-SBA-15 for the CO2
cycloaddition as described in Sect. 2.2.2, also examined the same catalysts for the
carbamate synthesis [23, 24]. From various aliphatic and aromatic amines, their
carbamates were produced with high selectivity under mild conditions (at 80  C and
0.34 MPa of CO2) in the absence of solvent. FTIR spectra of the catalysts measured
again revealed that the product yield was well correlated with the peak intensity of

Scheme 2.8 Synthesis of O


cyclic carbonate from CO2
HO OH
and glycol + CO2 O O + H2O
R
R

O
+ HX
RNH2 + CO2 + R⬘X RNH OR⬘

Scheme 2.9 Synthesis of carbamate from primary amine, CO2, and alkyl halide
2 Direct Transformation of Carbon Dioxide to Value-Added Products. . . 49

H
N O
NH2
Base R⬘
R⬘ + CO2
NH
CN
O

Scheme 2.10 Synthesis of quinazoline-2,4(1H,3H )-diones from CO2 and 2-aminobenzonitriles

Scheme 2.11 Oxidation of OH


benzene to phenol with CO2
+ CO2 + CO

surface carbamate species. The coexistence of Ti sites and base sites was again
suggested to be significant for the reaction. Adenine-modified Ti-SBA-15, which
was deactivated by recycling in the absence of solvent, could be recycled for
the carbamate synthesis without loss of activity even in the absence of solvent.
These catalysts would be of promise for realizing carbamate synthesis processes
from CO2 using solid catalysts.
The reactions of CO2 and aminobenzonitriles produce rather complex com-
pounds of quinazoline-2,4(1H,3H)-diones (Scheme 2.10), which are important
because of their biological activity and widely used as key structures in medical
drugs. Patil et al. employed a series of MgO/ZrO2 catalysts for this reaction
[54]. When 2-aminobenzonitrile was used as the substrate, pure MgO and ZrO2
gave the product yields of 44 % and 18 %, respectively, at 130  C and 3.5 MPa of
CO2 (at room temperature) for 12 h in DMF solvent. MgO/ZrO2 catalysts (MgO
loading 5–15 %) gave higher product yields and an almost quantitative yield was
obtained with 10 % MgO/ZrO2. MgO/SiO2 showed lower activity. FTIR of the
catalysts and CO2 TPD measurements suggested some interaction between MgO
and ZrO2. The most active catalyst gave good to excellent product yields from a
broad range of aminobenzonitriles. The selection of the solvent was significant. The
product yield obtained with DMSO was almost the same with DMF, while other
polar and nonpolar solvents were ineffective (0–10 % yields). The catalyst was
recyclable without loss of activity. The same group also reported the use of
inorganic base of Cs2CO3 and of organic one of 1-butyl-3-methyl imidazolium
hydroxide [55, 56]. The catalytic performance of the best MgO/ZrO2 catalyst was
comparable with those bases.
In all the above-mentioned reactions, CO2 was used as the source of carbonyl
group. CO2 can be an oxidant for the direct synthesis of phenol from benzene
(Scheme 2.11). Goettmann et al. first reported that graphitic carbon nitride C3N4
can catalyze this reaction in the co-presence of organic bases [57]. Lately, Ansari
et al. studied the influence of starting compounds to prepare C3N4 on the activity
[58]. In these studies, a high reaction temperature (150  C) and long reaction time
(10–20 h) were required to get reasonable product yields. So, further improvement
of the catalyst activity is still required to synthesize phenol in large scales.
50 S.-i. Fujita et al.

The by-product of CO can be used for other synthetic reactions, being an advantage
of this reaction. Furthermore, the current industrial phenol manufacturing processes
(the cumene process) suffer from the formation of large amounts of acetone as a
by-product. Considering these, the oxidation of benzene with CO2 would be a new
and highly desirable route for utilization of CO2.

2.5 Concluding Remarks

This chapter has reviewed the utilization of heterogeneous catalysts for the
transformation of CO2 to valuable chemicals. The reactions described are the CO2
cycloaddition to epoxide producing cyclic carbonate (Scheme 2.1), the synthesis of
DMC from CO2 and methanol (Scheme 2.6), and other several related reactions
(Schemes 2.8, 2.9, 2.10, and 2.11). These reactions are benign synthesis ways to
avoid the use of toxic chemicals of phosgene, CO, isocyanate, etc.
Inorganic heterogeneous catalysts of a few metal oxides, an alkali-loaded
zeolite, and smectites are effective for the CO2 cycloaddition in the absence of
any organic solvent. Furthermore, their activities are enhanced by the co-presence
of basic organic solvents. This enhancement is ascribed to the cooperative activa-
tion of the epoxide ring by the metal center and the base (Scheme 2.2). Synergistic
effect is observed on immobilization of organic bases including imidazolium
halides on inorganic solids having surface hydroxyl groups. This is also explained
by a similar cooperative activation mechanism (Scheme 2.4). Highly active het-
erogeneous bifunctional catalyst systems can also be realized by modification
(treatment with zinc halide or substituting with alkyl diol) of imidazolium halides
immobilized on polymers (Scheme 2.5). One of ZIFs, which are a new class of
porous materials, can be used for the CO2 cycloaddition.
DMC can be produced from CO2 and methanol in the presence of potassium
carbonate as a catalyst and methyl iodide as a promoter; however, the catalyst is
deactivated by transformation to potassium iodide, and methyl iodide acts as a
reactant rather than a promoter (Scheme 2.6). Even in the absence of methyl iodide,
ZrO2-based catalysts can selectively catalyze the DMC synthesis, and the removal
of H2O from the reaction system is essential to get high DMC yields because of the
reaction equilibrium limitation.
For the cyclic carbonate synthesis from CO2 and glycol, CeO2–ZrO2 is selective,
but it can give only low product yields because of the reaction equilibrium limitation.
Organic bases (tetraalkyl ammonium halides, adenine) immobilized on zeolite or
Ti-SBA-15 can give excellent product yields for the synthesis of carbamates from
CO2, primary amines, and alkyl halides. Also for this reaction, the coexistence of
acid Ti sites and base sites is suggested to be an important factor. MgO/ZrO2 is
effective for the synthesis of quinazoline-2,4(1H,3H)-diones from CO2 and
aminobenzonitriles. Graphitic carbon nitride can catalyze the oxidation of benzene
to phenol with CO2 that would be a new and highly desirable route for the
utilization of CO2.
2 Direct Transformation of Carbon Dioxide to Value-Added Products. . . 51

References

1. Edwards JH (1995) Potential sources of CO2 and options for its large–scale utilization now and
in future. Catal Today 23:59–66
2. Aresta M, Tomasi I (1997) Carbon dioxide utilization in the chemical industry. Energy
Convers Manage 38(Suppl):S373–S378
3. Arakawa H, Aresta M, Armor JM et al (2001) Catalysis research of relevance to carbon
management: progress, challenges and opportunities. Chem Rev 101:953–996
4. Zevenhoven R, Eloneva S, Teir S (2006) Chemical Fixation of CO2 in carbonates: routes to
valuable products and long–term storage. Catal Today 115:73–79
5. Ono Y (1997) Catalysis in the production and reactions of dimethyl carbonate, an environmentally
benign building block. Appl Catal A Gen 155:133–166
6. Bigi F, Maggi R, Sartori G (2000) Selected synthesis of ureas through phosgene substitutes.
Green Chem 2:140–148
7. Sakakura T, Choi JC, Yasuda H (2007) Transformation of carbon dioxide. Chem Rev
107:2365–2387
8. Sakakura T, Kohno K (2009) The synthesis of organic carbonates from carbon dioxide. Chem
Commun, pp 1312–1330
9. Shaikh AA, Sivaram S (1996) Organic carbonates. Chem Rev 96:951–976
10. Yano T, Matsui H, Koike T et al (1997) Magnesium oxide-catalysed reaction of carbon dioxide
with an epoxide with retention of stereo chemistry. Chem Commun, pp 1129–1130
11. Yamaguchi K, Ebitani K, Yoshida T et al (1999) Mg-Al mixed oxides as highly active
acid–base catalysts for cycloaddition of carbon dioxide to epoxides. J Am Chem Soc
121:4526–4527
12. Bhanage BM, Fujita S, Ikushima Y, Arai M (2001) Synthesis of dimethyl carbonate and
glycols from carbon dioxide, epoxides and methanol using heterogeneous basic metal oxide
catalysts with high activity and selectivity. Appl Catal A Gen 219:259–266
13. Tu M, Davis RJ (2001) Cycloaddition of CO2 to epoxides over solid base catalysts. J Catal
199:85–91
14. Nishiyama Y, Arai M, Guo SL et al (1993) Catalytic properties of hectorite-like smectites
containing nickel. Appl Catal A Gen 95:171–181
15. Fujita S, Bhanage BM, Ikushima Y et al (2002) Chemical fixation of carbon dioxide to
propylene carbonate using smectite catalysts with high activity and selectivity. Catal Lett
79:95–98
16. Bhanage BM, Fujita S, Ikushima Y et al (2003) Synthesis of dimethyl carbonate and glycols from
carbon dioxide, epoxides and methanol using heterogeneous Mg containing smectite catalysts:
effect of reaction variables on activity and selectivity performance. Green Chem 5:71–75
17. Yasuda H, He LN, Sakakura T (2002) Cyclic carbonate synthesis from supercritical carbon
dioxide and epoxide over lanthanide oxychloride. J Catal 209:547–550
18. Arest M, Dibenedetto A, Gianfrate L et al (2003) Nb(V) compounds as epoxides carboxylation
catalysts: the role of the solvent. J Mol Catal A Chem 204–205:245–252
19. Srivastava R, Srinivas D, Ratnasamy P (2003) Synthesis of polycarbonate precursors over
titanosilicate molecular sieves. Catal Lett 91:133–139
20. Mori K, Mitani Y, Hara T et al (2005) A single-site hydroxyapatite-bound zinc catalyst
for highly efficient chemical fixation of carbon dioxide with epoxides. Chem Commun,
pp 3331–3333
21. Sun J, Fujita S, Arai M (2005) Development in the green synthesis of cyclic carbonate from
carbon dioxide using ionic liquids. J Organometal Chem 690:3490–3497
22. North M, Pasquale R, Young C (2010) Synthesis of cyclic carbonates from epoxides and CO2.
Green Chem 12:1514–1539
23. Srivastava R, Srinivas D, Ratnasamy P (2005) Zeolite–based organic–inorganic hybrid catalysts
for phosgene–free and solvent–free synthesis of cyclic carbonates and carbamates at mild
conditions utilizing CO2. Appl Catal A Gen 289:128–134
52 S.-i. Fujita et al.

24. Srivastava R, Srinivas D, Ratnasamy P (2005) CO2 activation and synthesis of cyclic carbonates
and alkyl/aryl carbamates over adenine-modified Ti-SBA-15 solid catalysts. J Catal 233:1–5
25. Takahasi T, Watanuki T, Kitazume S et al (2006) Synergistic hybrid catalyst for cyclic
carbonate synthesis: remarkable acceleration caused by immobilization of homogeneous
catalyst on silica. Chem Commun, pp 1664–1666
26. Sakai T, Tsutsumi Y, Ema T (2008) Highly active and robust organic–inorganic hybrid
catalyst for the synthesis of cyclic carbonates from carbon dioxide and epoxides. Green
Chem 10:337–341
27. Dai WL, Chen L, Yin SF et al (2010) 3-(2-hydroxyethyl)-1-propylimidazolium bromide
immobilized on SBA-15 as efficient catalyst for the synthesis of cyclic carbonates via the
coupling of carbon dioxide and epoxides. Catal Lett 135:295–304
28. Motokura K, Itagaki S, Iwasawa Y et al (2009) Silica-supported aminopyridinium halides for
catalytic transformations of epoxides to cyclic carbonates under atmospheric pressure of
carbon dioxide. Green Chem 11:1876–1880
29. Udayakumar S, Lee MK, Shim HL, Park DW (2009) Functionalization of organic ions on
hybrid MCM-41 for cycloaddition reaction: the effective conversion of carbon dioxide. Appl
Catal A Gen 365:88–95
30. Kim HS, Kim JJ, Kim H, Jang HG (2003) Imidazolium zinc tetrahalide-catalyzed coupling
reaction of CO2 and ethylene oxide or propylene oxide. J Catal 220:44–46
31. Palgunadi J, Kwon O-S, Lee H et al (2004) Ionic liquid-derived zinc tetrahalide complexes:
structure and application to the coupling reactions of alkylene oxides and CO2. Catal Today
98:511–514
32. Fujita S, Nishiura M, Arai M (2010) Synthesis of styrene carbonate from carbon dioxide and
styrene oxide with various zinc halide-based ionic liquids. Catal Lett 135:263–268
33. Qiao K, Ono F, Bao Q et al (2009) Efficient synthesis of styrene carbonate from CO2 and
styrene oxide using zinc catalysts immobilized on soluble imidazolium-styrene copolymers.
J Mol Catal A Chem 303:30–34
34. Watile RA, Desmukh KM, Dhake KP, Bhanage BM (2012) Efficient synthesis of cyclic
carbonate from carbon dioxide using polymer anchored diol functionalized ionic liquids as a
highly active heterogeneous catalyst. Catal Sci Technol 2:1051–1055
35. Huang X-C, Lin Y-Y, Zhang J-P, Chen X-M (2006) Ligand-directed strategy for zeolite-type
metal-organic frameworks: zinc(II) imidazolates with unusual zeolitic topologies. Angew
Chem Int Ed 45:1557–1559
36. Banerjee R, Furukawa H, Britt D et al (2009) Control of pore size and functionality
in isoreticular zeolitic imidazolate frameworks and their carbon dioxide selective capture
properties. J Am Chem Soc 131:3875–3877
37. Miralda CM, Macias EE, Zhu M et al (2011) Zeolitic imidazole freamwork-8 catalysts in the
conversion of CO2 to chloropropene carbonate. ACS Catal 2:180–183
38. Ono Y (1997) Dimethyl carbonate for environmentally benign reactions. Catal Today
35:15–25
39. Fang S, Fujimoto K (1996) Direct synthesis of dimethyl carbonate from carbon dioxide and
methanol catalyzed by base. Appl Catal A Gen 142:L1–L3
40. Fujita S, Bhanage BM, Ikushima Y, Arai M (2001) Synthesis of dimethyl carbonate from
carbon dioxide and methanol in the presence of methyl iodide and base catalysts under mild
conditions: effect of reaction conditions and reaction mechanism. Green Chem 3:87–91
41. Tomishige K, Sakaihori T, Ikeda Y, Fujimoto K (1999) A novel method of direct synthesis of
dimethyl carbonate from methanol and carbon dioxide catalyzed by zirconia. Catal Lett
58:225–229
42. Ikeda Y, Sakaihori T, Tomishige K, Fujimoto K (2000) Catalytic properties and structure of
zirconia catalysts for direct synthesis of dimethyl carbonate from methanol and carbon
dioxide. Catal Lett 66:59–62
43. Tomoshige K, Furusawa Y, Ikeda Y et al (2002) CeO2–ZrO2 solid solution catalyst for
selective synthesis of dimethyl carbonate from methanol and carbon dioxide. Catal Lett
76:71–74
2 Direct Transformation of Carbon Dioxide to Value-Added Products. . . 53

44. Jung KT, Bell AT (2001) An in situ infrared study of dimethyl carbonate synthesis from carbon
dioxide and methanol over zirconia. J Catal 204:339–347
45. Jung KT, Bell AT (2001) Effects of catalyst phase structure on the elementary processes
involved in the synthesis of dimethyl carbonate from methanol and carbon dioxide over
zirconia. Top Catal 20:97–105
46. Tomishige K, Ikeda Y, Sakaihori T, Fujimoto K (2000) Promoting effect of phosphoric acid on
zirconia catalysts in selective synthesis of dimethyl carbonate from methanol and carbon
dioxide. J Catal 195:355–362
47. Jiang C, Guo Y, Wang C et al (2003) Synthesis of dimethyl carbonate from methanol and
carbon dioxide in the presence of polyoxometalates under mild conditions. Appl Catal A Gen
256:203–212
48. Tomishige K, Kunimori K (2002) Catalytic and direct synthesis of dimethyl carbonate starting
from carbon dioxide using CeO2–ZrO2 solid solution heterogeneous catalyst: effect of H2O
removal from the reaction system. Appl Catal A Gen 237:103–109
49. Li Z, Qin Z (2007) Synthesis of diphenyl carbonate from phenol and carbon dioxide in carbon
tetrachloride with zinc halides as catalyst. J Mol Catal A Chem 264:255–259
50. Fan G, Fujita S, Zou B et al (2009) Synthesis of diphenyl carbonate from phenol and carbon
dioxide in the presence of carbon tetrachloride with zinc chloride. Catal Lett 133:280–287
51. Fan G, Wang Z, Zou B, Wang B (2011) Synthesis of diphenyl carbonate from compressed
carbon dioxide and phenol without use of organic solvent. Fuel Process Technol 92:1052–1055
52. Li Z, Su K (2007) The direct reaction between CO2 and phenol catalyzed by bifunctional
catalyst ZrO2. J Mol Catal A Chem 277:180–184
53. Tomishige K, Yasuda H, Yoshida Y et al (2004) Catalytic performance and properties of ceria
based catalysts for cyclic carbonate synthesis from glycol and carbon dioxide. Green Chem
6:206–214
54. Patil YP, Tambade PJ, Parghi KD et al (2009) Synthesis of quinazoline-2,4(1H,3H )-diones
from carbon dioxide and aminobenzonitriles using MgO/ZrO2 as a solid base catalyst. Catal
Lett 133:71–74
55. Patil YP, Tambade PJ, Jagtop SR, Bhange BM (2008) Cesium carbonate catalyzed efficient
synthesis of quinazoline-2,4(1H,3H )-diones using carbon dioxide and 2-aminobenzonitriles.
Green Chem Lett Rev 1:127–132
56. Patil YP, Tambade PJ, Desmukh KM et al (2009) Synthesis of quinazoline-2,4(1H,3H )-diones
from carbon dioxide and aminobenzonitriles using [Bmim]OH as a homogeneous recyclable
catalyst. Catal Today 148:355–360
57. Goettmann F, Thomas A, Antonietti M (2007) Metal-free activation of CO2 by mesoporous
graphitic carbon nitride. Angew Chem Int Ed 46:2717–2720
58. Ansari MB, Min B-H, Mo Y-H, Park S-E (2011) CO2 activation and promotional effect in the
oxidation of cyclic olefins over mesoporous carbon nitrides. Green Chem 13:1416–1421
Chapter 3
Indirect Utilisation of Carbon Dioxide
in Organic Synthesis for Valuable Chemicals

Rahul A. Watile, Bhalchandra M. Bhanage, Shin-ichiro Fujita,


and Masahiko Arai

3.1 Introduction

Carbon dioxide (CO2) has attracted much attention from the scientific community
because of its potential greenhouse effect as a result of its steadily increasing
concentrations in the atmosphere [1–4]. Moreover, CO2 is nontoxic, nonflammable,
abundant and economical C1 feedstock. Currently, carbon dioxide is used indus-
trially for the synthesis of urea, salicylic acid, inorganic carbonates, methanol,
cyclic carbonates and polycarbonates [2–7]. Major emphasis is to replace
phosgene-based routes by introducing CO2 into organic substrates for synthesis of
various chemicals [5].
However, in industry, around 105 million tons of CO2 are utilised for the
production of urea. On the other hand, cyclic carbonates (80,000 t), salicylic acid
(90,000 t) and polypropylene carbonate (70,000 t) are produced on industrial scale
[6, 8]. In addition to these commercial industrial processes using CO2, there are a
lot of interesting approaches for the application of CO2. The simplest kind of
reaction is the insertion of CO2, e.g. to obtain carbamates [6, 9, 10], carboxylated
allyl derivatives [11], acetic acid [12, 13], dialkyl carbonates [14–17], polypyrones
[18], lactones [19–21] and polyurethanes [22, 23]. Although a lot of research has
been done on CO2 fixation reaction, there is still a challenge in utilising the CO2 as
such in organic transformation due to the requirement of high reaction temperatures
(250–300  C) and high pressures (50–100 atm) as well as thermodynamics limita-
tions in such direct reaction [24–27]. The development of a practical catalytic

R.A. Watile • B.M. Bhanage (*)


Department of Chemistry, Institute of Chemical Technology, Mumbai 400019, India
e-mail: bm.bhanage@ictmumbai.edu.in
S.-i. Fujita • M. Arai
Division of Chemical Process Engineering, Faculty of Engineering, Hokkaido University,
Sapporo 060-8626, Japan
e-mail: marai@eng.hokudai.ac.jp

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 55
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_3,
© Springer-Verlag Berlin Heidelberg 2014
56 R.A. Watile et al.

O O
HO NH2 O
H
N O 2 MeOH
O O
MeO OMe
NH2

NH2

H H
N N

Scheme 3.1 Organic carbonate used as sources of carbon dioxide for the industrially important
chemicals

system for CO2 fixation reaction still remains a challenge, as high activation energy
barriers have to be overcome for the cleavage of the C–O bonds of CO2 [28].
Cyclic carbonates are valuable synthetic targets; since their commercialisation
in the mid-1950s, the coupling of CO2 to epoxide to cyclic carbonates is one of the
most promising reactions in this area of replacing existing poisonous phosgene-
based synthesis. Cyclic carbonates have interesting applications as excellent aprotic
polar solvents, electrolytes in secondary batteries, engineering plastics and as raw
materials in the synthesis of polycarbonates and polyurethanes [10, 29–31]. Recently,
it has been shown that organic carbonates, e.g. EC, PC and DMC, can be used as a
carbonylation reagent for several organic transformations such as producing
various important chemicals such as dimethyl carbonate (DMC) via transester-
ification of EC with methanol. Furthermore, these carbonates have been success-
fully applied to the carbonylation of bifunctional substrates, such as diamines, and
amino alcohols which have been converted into cyclic carbamates [32–34],
N,N-bis(hydroxyalkyl)ureas [35, 36] and cyclic ureas [37]. Detailed develop-
ments in these areas are discussed here (Scheme 3.1). Salicylic acid is a crucial
intermediate for the synthesis of Aspirin. More than 100 years ago, the
Kolbe–Schmitt reaction for the synthesis of salicylic acid was developed and
still it is a topic of research [6, 8].
CO2 hydrogenation is an example that produces various chemicals depending on
the catalysts used. Effective copper-based catalysts have been explored for the
methanol synthesis from CO2 [38]. Recently, the direct reduction of CO2 to
methanol at mild conditions [39] and the development of new highly active
catalysts for the production of formic acid and its derivatives by CO2 hydrogenation
are still desirable [40–44]. The hydrogenation of carbon dioxide to CO, methane or
methanol has been far less explored, as these reactions are unfavourable from a
3 Indirect Utilisation of Carbon Dioxide in Organic Synthesis. . . 57

thermodynamic point of view. The hydrogenation of CO2 into CH3OH has been
extensively investigated, and Cu-/Zn-based catalyst was found to be highly selec-
tive, but under relatively harsh reaction conditions (250  C, 50 atm) [45, 46]. There-
fore, the methanol production from CO2 by direct hydrogenation under mild
reaction conditions is still a great challenge for both academia and industry
[42, 47–49]. Recently Milstein and co-workers developed an indirect approach
from CO2 to methanol through homogeneous hydrogenation of carbonates or
carbamates using pincer-type Ru(II) catalysts, under mild conditions [50]. Especially,
the hydrogenation of DMC has established a bridge from CO2 to methanol [24], as
DMC can be produced from CO2 and methanol [51]. In the present chapter, we focus
our results on many of these reactions (Scheme 3.1). Relevant studies by other
research groups are also compared.

3.2 Use of Organic Carbonates as Carbonylation Reagent

In recent years, it has been observed that organic carbonates, e.g. dimethyl carbon-
ate (DMC) and ethylene carbonate (EC), are of significant interest, because these
compounds can be produced from CO2 and used as a carbonylation reagent for
several synthetic protocols. EC is produced in large scale from CO2 with ethylene
oxide. It can react with other compounds, producing several valuable chemicals
(Scheme 3.1).
Organic carbonates are used as sources of CO2 for the synthesis of industrially
important chemicals (Scheme 3.1). This chapter mainly focuses on the ideal alter-
native routes for the synthesis of DMC, 1,3-disubstituted urea and 2-oxazolidinones/
2-imidazolidinones via transesterification of EC with methanol, transamination of EC
with primary amine and transamination reaction of EC with diamines/β-amino
alcohols, respectively, which will be described in the following section. Milstein
et al. reported for the first time catalytic hydrogenation of organic carbonates to
methanol, their mild hydrogenation which can provide alternative and mild
approaches to the indirect hydrogenation of CO2. Detailed developments of indirect
CO2 fixation reactions will be described in the following section.

3.3 Synthesis of DMC via Transesterification of EC

DMC has drawn specific attention as a nontoxic, noncorrosive and environmentally


friendly building block for the production of polycarbonates and variety of
chemicals. DMC finds extensive applications as a solvent and an octane booster
in gasoline to meet oxygenate specifications and is used as a starting material for
organic synthesis via carbonylation and methylation as well as promising substitute
for poisonous phosgene and dimethyl sulphate [52, 53]. In addition, it is used as an
electrolyte in lithium batteries due to its high dielectric constant [54]. Traditionally,
58 R.A. Watile et al.

O
CO2 + 2MeOH + H2O
MeO OMe

Scheme 3.2 Synthesis of DMC from CO2

O
O HO
O O + 2MeOH +
MeO OMe
OH

Scheme 3.3 Synthesis of DMC via transesterification of cyclic carbonate with methanol

DMC is synthesised by oxidative carbonylation of methanol (non-phosgene route)


or by phosgenation of methanol. Both routes involve the use of poisonous
and corrosive gases of chlorine, phosgene and carbon monoxide. Therefore, the
development of green and efficient route for DMC synthesis is till desirable.
From the green chemistry point of view, direct synthesis of DMC from CO2 is
one of the promising reactions for the chemical fixation of CO2 (Scheme 3.2)
[2, 55], which is more attractive compared with other commercial processes
including methanolysis of phosgene, carbon monoxide and carbonylation of meth-
anol using toxic, corrosive, flammable and explosive gases. Various catalysts are
reported for the direct synthesis of DMC from methanol and carbon dioxide,
including organometallic compounds [56], CeO2–ZrO2 [57] and H3PW12O40/
ZrO2 [58]. Among these catalysts, CeO2–ZrO2 showed a significant catalytic
activity for direct synthesis of DMC from methanol and CO2 due to its acid–base
bifunctional property. Although several reports exist for the direct DMC synthesis,
usually the reaction conditions employed are not mild, i.e. high reaction temperatures,
long reaction times and high CO2 pressures are needed. So the development of more
efficient and practical route for DMC synthesis is still desired.
DMC can also be synthesised via one-pot reaction directly from epoxides, CO2
and methanol [59]; however, undesired side reaction such as ring opening of
epoxides by methanol led to relatively lower DMC selectivity. The more promising
approach for synthesis of DMC based on transesterification of cyclic carbonates
with methanol comes forward (Scheme 3.3). The process for the synthesis of DMC
from the transesterification of EC with methanol was developed by Texaco Chem-
ical [60] (Scheme 3.3). Recently, Zhang et al. reported carboxylic functionalised
imidazolium ionic liquid as a catalyst for transesterification of cyclic carbonates
with methanol to DMC [61]. Ju et al. reported that DMC was prepared from
methanol and EC in the presence of 1-alkyl-3-methylimidazolium ionic liquids at
160  C and 2 MPa CO2 [62]. Yang et al. reported DMC synthesis (with 81 % yield)
via transesterification of EC with methanol, and it was achieved by employing the
Lewis basic ILs derived from DABCO as efficient and recyclable catalysts
[63]. Most of the reported heterogeneous catalysts require harsh reaction conditions
like high temperature and pressure to obtain even a moderate yield of DMC
3 Indirect Utilisation of Carbon Dioxide in Organic Synthesis. . . 59

[64–67]. Feng et al. used amino-functionalised MCM-41 as a heterogeneous


catalyst to obtain 44 mol% DMC yield at 160  C [68]. Kinfton and Duranleau
reported organic phosphine/amine catalysts supported on partially cross-linked
polymer for the synthesis of DMC which gave 58 % conversion of EC [69]. Tatsumi
et al. reported K-TS-1 as an effective catalyst with DMC yield of 57 %
[64]. Watanabe and Tatsumi demonstrated the use of Mg–Al hydrotalcite materials
as an active catalyst for transesterification reaction and they observed 70 % EC
conversion with 58 % DMC yield. The major drawback of all mentioned hetero-
geneous catalyst systems is that they gave lower yields of DMC. Although great
advances have been made, most of the catalytic system suffer from low catalytic
activity, high reaction temperature and the use of expensive metallic compounds
that are often required. There is a need to develop a suitable catalytic system for the
synthesis of DMC via transesterification of cyclic carbonate. After considering all
the merits and the demerits of the reported methods, one of our contribution in the
investigation of transesterification of EC with methanol for synthesis of DMC is
using poly-4-vinyl pyridine (PVP) as a novel, homogeneous recyclable base cata-
lyst [70] (Scheme 3.3). Basic properties of the catalyst was examined by temper-
ature programmed desorption (TPD) of adsorbed CO2 using a conventional flow
reactor. Basicity of PVP is responsible for high activity and selectivity for the
synthesis of DMC. Optimisation of the reaction parameters were investigated for
the transesterification reaction of EC with methanol in presence of PVP which
results in the formation of DMC and EG. The higher conversion of EC (96 %) and
higher yield of DMC (82 %) and EG (75 %) was observed at 140  C reaction
temperature, 20 wt-% of PVP loading and for reaction time of 4 h. It has been
observed that the CO2 pressure does not have any influence on the conversion
of EC and yield of DMC/EG. It was found that the catalyst could be active
up to three consecutive recycles without significant loss in its catalytic activity
and selectivity.

3.4 Synthesis of 2-Oxazolidinones/2-Imidiazolidinones


via Transesterification of EC with β-Amino
Alcohols/1,2-Diamines

Synthesis of five-member oxazolidinones and imidiazolidinones, possessing wide


application as intermediates, chiral auxiliaries and building blocks for biologically
active pharmaceutical agents, in organic synthesis are desirable [71, 72]. They show
good antibacterial properties, hence widely used in pharmaceuticals, cosmetics,
pesticides and so on. Chiral 2-oxazolidinones have been used as chiral auxiliaries in
a wide range of asymmetric synthesis [73–75]. 2-Imidiazolidinones have recently
attracted much attention due to their diverse applications as intermediates for
biologically active pharmaceutical molecules, such as the HIV protease inhibitors,
DMP 323 and DMP 450 [76]. The developments of greener and efficient route for
60 R.A. Watile et al.

O H2N OH

O O +
H2N NH2

H2N OH O O H2N OH
CO + O2 + CO2 +
HN HN O
H2N NH2 H2N NH2

O H2N OH
+
H2N NH2
H2N NH2

Scheme 3.4 Various routes for the synthesis of 2-imidiazolidinones and 2-oxazolidinones

synthesis of 2-oxazolidinones/2-imidiazolidinones have attracted much more attention


in recent years.
Usually 2-imidiazolidinones and 2-oxazolidinones are synthesised by carbonyl-
ation of 1,2-diamines and β-amino alcohols with several reagents like phosgene
[77], urea, dialkyl carbonates [78, 79] or the mixture CO/O2 via oxidative carbon-
ylation (Scheme 3.4). However, the mentioned synthesis routes are often associated
with the use of high pressure, high temperature, formation of side product like
polyurea and risk associated with explosion hazards using CO/O2 mixtures [80, 81].
Direct addition of CO2 with alkylene diamines and amino alcohols to produce
2-imidiazolidinones and 2-oxazolidinones is the most desired route (Scheme 3.4).
However, the synthesis is often associated with the use of high pressure and high
temperature and formation of side product [82]. Hence, development of an alter-
native for improving the yield as well as selectivity is more desirable. The
2-imidiazolidinones/2-oxazolidinones could be synthesised using transamination
reaction of EC with diamines/β-amino alcohols in the presence of a base catalyst
(Scheme 3.5) [83]. This reaction can be considered as indirect CO2 fixation
reaction, since EC is synthesised with a quantitative yield by the cycloaddition of
CO2 to epoxides [84, 85]. Recently, Xiao et al. reported the synthesis of 2-imidiazo-
lidinones/2-oxazolidinones from EC with 1,2-diamines/β-amino alcohols in the
presence of a homogenous catalyst [83]. However, it suffers from the major
drawback of catalyst–product separation. We investigated transamination reaction
of EC with diamines/β-amino alcohols using heterogeneous base catalysts such as
basic metal oxides for the first time, an indirect CO2 fixation reaction [86]
(Scheme 3.5). Several commercially available metal oxides such as MgO, CaO,
ZnO, ZrO2, CeO2 and La2O3 were screened for the reaction. The basic properties of
3 Indirect Utilisation of Carbon Dioxide in Organic Synthesis. . . 61

H2N OH
O

R2 HO OH
HN O +
O
R1
R2
O O
O
R1 R1
HN NH +
H2N NH2 HO OH
R3
R3

Scheme 3.5 Synthesis of 2-oxazolidinones/2-imidiazolidinones from cyclic carbonates and


β-amino alcohols/1, 2-diamines

the catalysts were examined by temperature programmed desorption (TPD) of


adsorbed CO2 using a conventional flow reactor. It was observed that among the
various metal oxide catalysts screened, only MgO gave excellent yield of
2-oxazolidinones (83 %) after 6 h. CaO and CeO2 were also found to give good
yield of 2-oxazolidinones 68 % and 72 %, respectively. The effects of various
reaction parameters like reaction time, reaction temperature, solvent effect and
catalyst loading on the reaction system were also examined. It was observed that the
yields of the 2-oxazolidinones were found to increase to 83 % at a temperature of
80  C after the reaction time of 6 h in the ethanol solvent. It was observed that the
catalyst can be reused for four conjugative recycle runs without any significant loss
in the catalytic activity.

3.5 Synthesis of 1,3-Disubstituted Symmetrical/


Unsymmetrical Urea

1,3-Disubstituted ureas are very important classes of nitrogen-containing com-


pounds, having a wide range of application in dyes for cellulose fibre, as antioxi-
dants in gasoline, as corrosion inhibitors and intermediates for carbamate synthesis
[87] as well as showing pharmacological and physiological activities. Recently,
special interest in these compounds has arisen, since substituted ureas having amino
acid group have been used for brain cancer treatment and were proven to have a
marked inhibitory effect on HIV protease enzyme [88]. Traditionally, N,N0 -alkyl
aryl ureas and N,N0 -dialkyl ureas are prepared by the nucleophilic addition of aryl
amines and alkyl amines to phosgene or isocyanates. However, phosgene or iso-
cyanates are highly toxic and hazardous reagents. Therefore, these routes are not
considered as eco-friendly routes for the urea synthesis. Oxidative carbonylation of
amine with CO is one of the phosgene-free alternative routes for the urea synthesis
62 R.A. Watile et al.

O
2R NH2 + CO + O2 R R
N N + H2O
H H

Scheme 3.6 Synthesis of substituted ureas via oxidative carbonylation of amines with CO

O
2R NH2 + CO2 R R + H O
N N 2
H H

Scheme 3.7 Synthesis of substituted ureas from CO2 and amines

O O O
CsCO3
R R
2R NH2 + O O R
N O
OH N N + HO OH
H H H

Scheme 3.8 Synthesis of symmetrical 1,3-disubstituted urea

(Scheme 3.6). Various homogeneous metal catalysts, including Au [89], Pd [90]


and W [91], have been reported for the synthesis of 1, 3-disubstituted ureas via
oxidative carbonylation. Deng et al. demonstrated the use of ZrO2–SO2 supported
palladium catalyst for the synthesis of 1,3-disubstituted ureas [92]. This method
suffers from the drawback associated with the tremendous toxicity of CO as well as
the risk associated with the use of explosive mixture of CO and O2. Hence, an idea
of direct synthesis of 1,3-disubstituted ureas from CO2 and amines came forward
(Scheme 3.7) [93–98].
Different catalysts have been reported for the synthesis of 1, 3-disubstituted urea
from CO2, such as RuCl33H2O/Bu3P and Ph3SbO/P4S10 [99]. Water is formed
during the course of reaction. The dehydrating agents such as POCl3, propargyl
alcohols [96], diorganophosphites [98], Me3NSO3 [97] and carbodiimide [95] are
required to increase the yield of the target product. Vos and co-workers achieved
the synthesis of symmetrical and unsymmetrical urea in good yields from CO2 and
amines using Cs+ base catalysts and N-methylpyrrolidone as the solvent without a
dehydrating agent [93]. With increasing demands for safer and cleaner processes,
the hazardous phosgene process has to be essentially replaced by more environ-
mentally friendly processes; e.g. the transesterification of EC with amine
(Scheme 3.8) shows promise as a phosgene-free route.
From the literature, we observed that EC-based CO2 fixation reactions have
received much more attention to organic chemistry because these are considered
to be indirect CO2 fixation reactions. EC represents an effective carbonylating
reagent [100] (Scheme 3.8). In 1998, Yasuoka et al. patented a synthesis of
1,3-disubstituted urea via transamination reaction of EC with an excess of
butylamine in methanolic NaOMe solution (28 %) and in an autoclave at 100  C.
It gave the product 1,3-dibutylurea in 85 % yield [101]. There are very few reports
3 Indirect Utilisation of Carbon Dioxide in Organic Synthesis. . . 63

on transamination reaction between EC and amines. Arai and co-workers reported


the synthesis of both symmetrical and unsymmetrical 1,3-disubstituted urea using
alkali metal oxides as heterogeneous base catalyst for the first time [102,
103]. Recently, our group reported the synthesis of 1,3-disubstituted symmetrical/
unsymmetrical urea via transamination of in situ generated 2-hydorxycarbamate
(Scheme 3.8) [104] using Cs2CO3 as heterogeneous base catalysts. Various com-
mercially available basic catalysts like alkali metal carbonates/hydroxides/halides
were screened for the reaction. It can be seen that among the various base catalysts
screened, only Cs2CO3 was the most successful base delivering 1,3-dibutylurea in
quantitative yield. Other carbonate bases were not as effective for the transamina-
tion reaction. The high catalytic reactivity of Cs2CO3 is attributed to the unprece-
dented ‘cesium effect’ [104, 105] and due to the softness of Cs cation, which makes
Cs2CO3 highly soluble in reaction mixture. Hence, we thought this methodology
therefore will contribute to existing cesium base-promoted transformations,
whereas alkali metal halide was found to be inactive for the synthesis of
1,3-dibutylurea. Hence, in order to optimise reaction conditions, we examined
the effect of temperature and reaction time on the 1,3-disubstituted urea synthesis.
It has been observed that at 90  C, Cs2CO3 gives 91 % yield of the 1,3-dibutylurea.
The reaction was carried out using various aliphatic primary, secondary and
heterocyclic as well as aromatic amines. It was observed that aliphatic primary
amines could efficiently react with EC in the presence of Cs2CO3, which leads to
the formation of corresponding symmetrical 1,3-disubstituted ureas. Unfortunately,
aromatic amines and secondary amine are unreactive under the given optimised
reaction condition.

3.6 Synthesis of Unsymmetrical 1,3-Disubstituted Ureas

On the basis of the results obtained for the synthesis of symmetrical


1,3-disubstituted urea, we next directed our attention towards the synthesis of
unsymmetrical 1,3-disubstituted urea (Scheme 3.9). Recently, Saliu et al. reported
the solventless method for the synthesis of unsymmetrical ureas via transamination
reaction between EC and amines using 1,5,7-Triazabicyclo[4.4.0]dec-5-ene (TBD)
and thioureas as organocatalysts for this reaction [106]. We developed a two-step
procedure for the synthesis of unsymmetrical ureas (Scheme 3.9). In the first step a
primary amine reacts with EC to give the corresponding 2-hydroxyethylcarbamate
at 70  C for 1 h in the absence of a catalyst [107]. In the second step, the
2-hydroxyethylcarbamate reacts in the presence of Cs2CO3 base catalyst with
another primary amine or with a secondary amine to give disubstituted or trisub-
stituted ureas, respectively. Although this method is fast and providing high yield,
the formation of symmetrical ureas as side products is its major drawback, and
column chromatography is needed to obtain the desired products in high purity.
Hence, development of more efficient, easily prepared and one-pot protocol for the
synthesis of unsymmetrical 1,3-disubstituted urea is still highly desired.
64 R.A. Watile et al.

O O O
2
R1 R 2 + R1 R1 +R N R2
N N N N N
NH2 H H
O 2 H H H H
R
O O O R1 TBD
HO
R1 NH2 R1 OH N
N O C
H O OH3
R
NH O
O
R4 R1 1
R3 + R N N R
1
TBD N N
H R 4 H H
R1 = cyclohexyl, R3 = R4 = ethyl

Scheme 3.9 Synthesis of unsymmetrical 1,3-disubstituted urea via transamination of EC with


amines

O
2 MeOH + CO2 + H2O
MeO OMe

Scheme 3.10 Synthesis of DMC from carbon dioxide

O PNN Ru(II)
pincer
+ 3H2 3MeOH
MeO OMe complexes

Scheme 3.11 Hydrogenation of organic carbonates (such as DMC) to methanol

3.7 Efficient Hydrogenation of Organic Carbonates:


Alternative Routes to Methanol Based on CO2

The industrial production of methanol by direct hydrogenation of CO2 under mild


conditions is an attractive goal, but a practical catalytic process has not yet been
developed. Methanol is an important fuel and synthetic building block, industrially
produced from syngas at high temperatures (250–300  C) and high pressures
(50–100 atm) using a copper oxide-based catalyst [25]. Hydrogenation of methyl
formate to methanol is known [43, 108–111]. However, this method suffers from
drawback such as the formation of CO by-product from the decomposition of
methyl formate, making it less attractive for large-scale application. In contrast,
the carbonylation of methanol to DMC [112] or oxidative carbonylation to DMC
and the synthesis of DMC from CO2 are known and well investigated [4]
(Scheme 3.10).
To the best of our knowledge, the hydrogenation of organic carbonates (such
as dimethyl carbonate) to methanol remains unknown [109, 113]. Milstein et al.
reported the first examples of catalytic hydrogenation of dimethyl carbonate to
methanol [114] (Scheme 3.11). They used PNN Ru(II) pincer complexes derived
from pyridine- and bipyridine-based tridentate ligands. Thus, catalytic hydrogenation
3 Indirect Utilisation of Carbon Dioxide in Organic Synthesis. . . 65

Catalyst, H2,T °C
CaCO3 or MgCO3 Hydrocarbon (C1-C3)

Scheme 3.12 Catalytic hydrogenation of calcium carbonate to hydrocarbons

of the organic carbonates to methanol under mild conditions could provide an indirect
method to obtain methanol from CO2. These new reactions provide interesting
alternatives for the mild hydrogenation of CO2 to methanol representing the ultimate
‘green’ reactions, which is a timely topic in the context of the ‘methanol economy’
[24]. Further developments of catalyst and process design will still be needed for this
reaction to be industrialised.

3.8 Direct Conversion of Calcium Carbonate


to Hydrocarbons

One of the central challenges in the twenty-first century is the sufficient and
sustainable supply of energy. At present, we derive more than 80 % of our energy
from fossil fuels [115]. Inorganic carbonates constitute nontoxic, nonflammable,
inexpensive natural source of carbon. Therefore, the objective of investigating the
direct conversion of these carbonates into hydrocarbons is of considerable impor-
tance (Scheme 3.12).
Replacing gaseous CO2 with solid inorganic carbonates which are abundantly
present in nature, such as calcite (CaCO3) and dolomite (MgCO3), has many
inherent advantages. In the decomposition of the carbonates, the recarbonisation
process can be used to keep the resources renewable for the production of fuels and
chemicals. Here, we focused on inorganic carbonates, instead of gaseous CO2, as a
source of carbon to form methane and studied the efficiency of the conversion. CNR
Rao and co-workers reported for the first time the catalytic hydrogenation of
inorganic carbonates to methane using Co/CoO/CaO as mixed metal oxide-based
catalytic protocol (Scheme 3.12) [116]. This route can be considered as an indirect
hydrogenation of CO2 as they can be rapidly replenished by the carbonation process
from CO2. Recently, the same group reported the hydrogenation of CaCO3 to form
C1–C3 hydrocarbons, in the presence of iron-based catalysts [117]. Such a study
would also be relevant to convert carbonaceous inorganic rocks to organics, which
can be of importance in the future. The research on catalytic hydrogenations of
inorganic carbonates must be developed further.

3.9 Conclusions and Future Trends

This chapter addresses recent contribution to the development of innovative and


environmentally friendly chemical processes based on CO2. The utilisation of CO2
in organic synthesis is important with regard to securing carbon resources as well as
66 R.A. Watile et al.

mitigating the greenhouse effect. It is noteworthy that cyclic carbonate synthesis


has already been an industrialised process. However, the development of new
catalysts for this reaction continues to be a fruitful area of research, and recent
advance such as alkali metal halide-supported liquid phase catalyst (SLPC) and
silica-supported catalysts have the potential to allow this industrial process to be in
actual practice. The synthesis of DMC and direct hydrogenation of CO2 to methanol
are attractive goals in organic synthesis, but a practical catalytic process has not yet
been developed because of the reaction equilibrium and chemical inertness of CO2.
The preferred alternatives are the transesterification of EC with methanol, trans-
amination reaction of EC with diamines/β-amino alcohols and catalytic hydroge-
nation of DMC, an indirect route to methanol from CO2. Further significant
developments in catalysis and process design will occur in the next few years;
thus, CO2 can be used for synthesis of industrially important chemicals which can
be used to help to limit global CO2 emissions.
Additionally, the future of CO2 lies in its viable conversion to CO via partial
reduction using renewable source of hydrogen, and once we can get CO at ambient
conditions, it opens new avenues for CO-based well-known chemistry. This reac-
tion is overlooked so far, and research in this direction is required.

References

1. Jessop PG, Ikariya T, Noyori R (1995) Homogeneous catalysis in supercritical fluids. Science
269:1065–1069
2. Aresta M, Quaranta E (1997) Carbon dioxide: a substitute for phosgene. Chemtech 27:32–40
3. Shaikh AA, Shivaram S (1995) Organic carbonates. Chem Rev 96:951–957
4. Sakakura S, Choi JC, Yasuda H (2007) Transformation of carbon dioxide. Chem Rev
107:2365–2387
5. Pierre B, Dominque M, Dominique N (1988) Reaction of carbon dioxide with carbon -carbon
bond formation catalyzed by transition-metal complexes. Chem Rev 88:747–764
6. Aresta M (2003) Carbon dioxide recovery and utilization. (A summary report of the EU
Project BRITE-EURAM 1998 BBRT-CT98-5089.) Kluwer Academic Publishers, Dordrecht,
407 pp
7. Lindsey AS, Jeskey H (1957) The Kolbe-Schmitt reaction. Chem Rev 57:583–620
8. Gibson DH (1996) The organometallic chemistry of carbon dioxide. Chem Rev
96:2063–2096
9. Chaturvedi D, Ray S (2006) Versatile use of carbon dioxide in the synthesis of carbamates.
Monatsh Chem 137:127–145
10. Rohr M, Geyer C, Wandeler R, Schneider MS, Murphy EF, Baiker A (2001) Solvent-free
ruthenium-catalysed vinylcarbamate synthesis from phenylacetylene and diethylamine in
supercritical carbon dioxide. Green Chem 3:123–125
11. Shi M, Nicholas KM (1997) Palladium-catalyzed carboxylation of allyl stannanes. J Am
Chem Soc 119:5057–5058
12. Shi DX, Feng YQ, Zhong SH (2004) Photocatalytic conversion of CH4 and CO2 to oxygen-
ated compounds over Cu/CdS–TiO2/SiO2 catalyst. Catal Today 98:505–509
13. Wilcox EM, Roberts GW, Spivey JJ (2003) Direct catalytic formation of acetic acid from
CO2 and methane. Catal Today 88:83–90
3 Indirect Utilisation of Carbon Dioxide in Organic Synthesis. . . 67

14. Sakakura T, Choi J, Saito CY, Sako T (2000) Synthesis of dimethyl carbonate from carbon
dioxide: catalysis and mechanism. Polyhedron 19:573–576
15. Iwakabe K, Nakaiwa M, Sakakura T, Choi JC, Yasuda H, Takahashi T, Ooshima Y (2005)
Reaction rate of the production of dimethyl carbonate directly from the supercritical CO2 and
methanol. J Chem Eng Jpn 38:1020–1024
16. Aresta M, Dibenedetto A, Fracchiolla E, Giannoccaro P, Pastore C, Papai I, Schubert G
(2005) Mechanism of formation of organic carbonates from aliphatic alcohols and carbon
dioxide under mild conditions promoted by carbodiimides. J Org Chem 70:6177–6186
17. Tomishige K, Sakaihori T, Ikeda Y, Fujimoto K (1999) A novel method of direct synthesis of
dimethyl carbonate from methanol and carbon dioxide catalyzed by zirconia. Catal Lett
58:225–229
18. Tsuda T, Maruta K, Kitaike Y (1992) Nickel (0)-catalyzed alternating copolymerization of
carbon dioxide with diynes to poly(2- pyrones). J Am Chem Soc 114:1498–1499
19. Behr A, Heite M (2000) Telomerization of carbon dioxide and 1, 3-butadiene: process
development in a miniplant. Chem Eng Technol 23:952–955
20. Schulz PS, Walter O, Dinjus E (2005) Facile synthesis of a tricyclohexylphosphine-stabilized
η3-allyl-carboxylato Ni(II) complex and its relevance in electrochemical butadiene carbon
dioxide coupling. Appl Organomet Chem 19:1176–1179
21. Takimoto M, Mori M (2002) Novel catalytic CO2 incorporation reaction: nickel-catalyzed
regio- and stereoselective ring-closing carboxylation of bis-1,3-dienes. J Am Chem Soc
124:10008–10009
22. McGhee W, Riley D, Christ K, Pan Y, Parnas B (1995) Carbon dioxide as a phosgene
replacement: synthesis and mechanistic studies of urethanes from amines, CO2 and alkyl
chlorides. J Org Chem 60:2820–2830
23. Ochiai B, Inoue S, Endo T (2005) One-pot non-isocyanate synthesis of polyurethanes from
bisepoxide, carbon dioxide and diamine. J Polym Sci Part A Polym Chem 43:6613–6618
24. Olah GA, Geoppert A, Surya Prakash GK (2009) Beyond oil and gas: the methanol economy,
2nd edn. Wiley-VCH, Weinheim
25. Bart JCJ, Sneeden JJF (1987) Copper-zinc oxide-alumina methanol catalysts revisited. Catal
Today 2:1–124
26. Weissermel K, Arpe HJ (2003) Industrial organic chemistry. Wiley-VCH, Weinheim,
pp 30–36
27. Hansen JB, Nielsen PEH (2008) In: Ertl G, Knçzinger H, Schüth F, Weitkamp J (eds)
Handbook of heterogeneous catalysis, vol 6, 3rd edn. Wiley-VCH, Weinheim, pp 2920–2949
28. Omae I (2006) Aspects of carbon dioxide utilization. Catal Today 115:33
29. Behr A (1988) Carbon dioxide as an alternative C1 synthetic unit: activation by transition-
metal complexes. Angew Chem Int Ed Engl 27:661–678
30. Elvers B, Hawkins S, Schulz G (eds) (1992) Ullmann’s encyclopedia of industrial
chemistry A, vol 21, 5th edn. VCH, Weinheim, p 207
31. Beckman EJ (1999) Making polymers from carbon dioxide. Science 283:946–947
32. Giannoccaro P, Dibenedetto A, Gargano M, Quaranta E, Aresta M (2008) Interaction of
palladium(II) complexes with amino-alcohols: synthesis of new amino-carbonyl complexes,
key intermediates to cyclic carbamates. Organometallics 27:967–975
33. Gabriele B, Salerno G, Brindisi D, Costa M, Chiusoli GP (2000) Synthesis of
2-oxazolidinones by direct palladium-catalyzed oxidative carbonylation of 2-amino-1-
alkanols. Org Lett 2:625–627
34. Gabriele B, Mancuso R, Salerno G, Costa DM (2003) An improved procedure for the
palladium-catalyzed oxidative carbonylation of β-amino alcohols to oxazolidin-2-ones.
J Org Chem 68:601–604
35. Giannoccaro P, Ferragina C, Gargano M, Quaranta E (2010) Pd-catalysed oxidative carbon-
ylation of amino alcohols to N, N0 -bis(hydroxyalkyl)ureas under mild conditions using
molecular oxygen as the oxidant. Appl Catal A Gen 375:78–84
36. Diaz DJ, Hylton KG, McElwee-White L (2006) Selective catalytic oxidative carbonylation of
amino alcohols to ureas. J Org Chem 71:734–738
68 R.A. Watile et al.

37. Qian F, McCusker JE, Zhang Y, Main AD, Chlebowski M, Kokka M, McElwee-White L
(2002) Catalytic oxidative carbonylation of primary and secondary diamines to cyclic ureas.
Optimization and substituent studies. J Org Chem 67:4086–4092
38. Ushikoshi K, Mori K, Watanabe T, Saito M (1998) A 50 kg/day class test plant for methanol
synthesis from CO2 and H2 Study. Surf Sci Catal 114:357–364
39. Das S, Mçller K, Junge K, Beller M (2011) Zinc-catalyzed chemoselective reduction of esters
to alcohols. Chem Eur J 17:7414
40. Darensbourg DJ, Ovalles C, Pala M (1983) Homogeneous catalysts for carbon dioxide/
hydrogen activation. Alkyl formate production using anionic ruthenium carbonyl clusters
as catalysts. J Am Chem Soc 105:5937–5939
41. Tsai JC, Nicholas KM (1992) Rhodium-catalyzed hydrogenation of carbon dioxide to formic
acid. J Am Chem Soc 114:5117–5124
42. Jessop PJ, Ikariya T, Noyori R (1995) Homogeneous hydrogenation of carbon dioxide. Chem
Rev 95:259–272
43. Jessop PJ, Hisiao Y, Ikariya T, Noyori R (1996) Homogeneous catalysis in supercritical
fluids: hydrogenation of supercritical carbon dioxide to formic acid, alkyl formates, and
formamides. J Am Chem Soc 1996(118):344–355
44. Tominaga K, Sasaki Y, Kawai M, Watanabe T, Saito M (1993) Ruthenium complex catalysed
hydrogenation of carbon dioxide to carbon monoxide, methanol and methane. J Chem Soc
Chem Commun 7:629
45. Boddien A, Gärtner F, Federsel C, Piras I, Junge H, Jackstell R, Beller M (2012) In: Ding K,
Dai L-X (eds) Organic chemistry: breakthroughs and perspectives. Wiley-VCH, Weinheim,
pp 683–722
46. Wang W, Wang SP, Ma XB, Gong JL (2011) Recent advances in catalytic hydrogenation of
carbon dioxide. Chem Soc Rev 40:3703
47. Marinus JM, Nico SW, Hiemstra H, Poetsch E, Casutt M (1995) Synthesis of D-(+)-biotin
through selective ring closure of N-acyliminium silyl enol ethers. Angew Chem
34–21:2391–2393
48. Leitner W (1995) Carbon dioxide as a raw material: the synthesis of formic acid and its
derivatives. Angew Chem Int Ed Engl 34:2207
49. Ma J, Sun N, Zhang X, Zhao N, Mao F, Wei W, Sun Y (2009) A short review of catalysis for
CO2 conversion. Catal Today 148:221
50. Balaraman E, Gunanathan C, Zhang J, Shimon LJW, Milstein D (2011) Efficient hydroge-
nation of organic carbonates, carbamates and formates indicates alternative routes to meth-
anol based on CO2 and CO. Nat Chem 3:609
51. Dixneuf PH (2011) Bifunctional catalysis: a bridge from CO2 to methanol. Nat Chem 3:578
52. Sato Y, Yamamoto T, Souma Y (2000) Poly(pyridine-2,5-diyl)-CuCl2 catalyst for synthesis
of dimethyl carbonate by oxidative carbonylation of methanol: catalytic activity and corro-
sion influence. Catal Lett 65:123–126
53. Pacheco MA, Marshall CL (1997) Review of dimethyl carbonate (DMC) manufacture and its
characteristics as a fuel additive. Energy Fuels 11:2–29
54. Wei T, Sun Y (2003) Synthesis of dimethyl carbonate by transesterification over CaO/carbon
composites. Green Chem 5:343
55. Jessop PG, Ikariya T, Noyori R (1999) Homogeneous catalysis in supercritical fluids. Chem
Rev 99:475
56. Taniguchi T, Ogasawara K (1993) Lipase–triethylamine-mediated dynamic transesteri-
fication of a tricyclic acyloin having a latent meso-structure: a new route to optically pure
oxodicyclopentadiene. Chem Commun 15:1399
57. Tomishige K, Furusawa Y, Ikeda Y, Asadullah M, Fujimoto K (2001) CeO2–ZrO2 solid
solution catalyst for selective synthesis of dimethyl carbonate from methanol and carbon
dioxide. Catal Lett 76:71
58. Jiang C, Guo Y, Wang C, Hu C, Wu Y, Wang E (2003) Synthesis of dimethyl carbonate from
methanol and carbon dioxide in the presence of polyoxometalates under mild conditions.
Appl Catal A Gen 256:203–212
3 Indirect Utilisation of Carbon Dioxide in Organic Synthesis. . . 69

59. Bhanage BM, Fujita S, Ikushima Y, Arai M (2001) Synthesis of dimethyl carbonate and
glycols from carbon dioxide, epoxides, and methanol using heterogeneous basic metal oxide
catalysts with high activity and selectivity. Appl Catal A Gen 219:259–266
60. Stinson S (1992) Edith M. Flanigen Wins Perkin Medal. Chem Eng News 10:25
61. Wang JQ, Sun J, Cheng WG, Shi CY, Dong K, Zhang XP, Zhang SJ (2012) Synthesis of
dimethyl carbonate catalyzed by carboxylic functionalized imidazolium salt via transester-
ification reaction. Catal Sci Technol 2:600–605
62. Ju HY, Manju MD, Park DW (2007) Performance of ionic liquid as catalysts in the
synthesis of dimethyl carbonate from ethylene carbonate and methanol. React Kinet
Catal Lett 90:3
63. Yang ZZ, He LN, Dou XY, Chanfreau S (2010) Dimethyl carbonate synthesis catalyzed by
DABCO-derived basic ionic liquids via transesterification of ethylene carbonate with meth-
anol. Tetrahedron Lett 51:2931–2934
64. Tatsumi T, Watanabe Y, Koyano KA (1996) Synthesis of dimethyl carbonate from ethylene
carbonate and methanol using TS-1 as solid base catalyst. Chem Commun 19:2281–2283
65. Sriniwas D, Srivasatava R, Ratnasamy P (2004) Transesterifications over titanosilicate
molecular sieves. Catal Today 96:127–133
66. Fujita S, Arai M (2005) Chemical fixation of carbon dioxide: synthesis of cyclic carbonate,
dimethyl carbonate, cyclic urea and cyclic urethane. J Japan Pet Inst 48:67
67. Li Y, Zhao XQ, Wang YJ (2004) Synthesis of dimethyl carbonate from propylene oxide,
carbon dioxide and methanol on KOH/4A molecular sieve catalyst. Chin J Cat 25:633
68. Feng SJ, Lu ZB, He R (2004) Tertiary amino group covalently bonded to MCM-41 silica as
heterogeneous catalyst for the continuous synthesis of dimethyl carbonate from methanol and
ethylene carbonate. Appl Catal A Gen 272:347–352
69. Knifton JF, Duranleau RG (1991) Ethylene glycol-dimethyl carbonate cogeneration. J Mol
Catal 67:389–399
70. Jagtap SR, Bhanushali MJ, Panda AG, Bhanage BM (2008) Synthesis of dimethyl carbonate
via transesterification of ethylene carbonate with methanol using poly-4-vinyl pyridine as a
novel base catalyst. Catal Commun 122:1928–1931
71. Dyen ME, Swern D (1967) 2-Oxazolidones. Chem Rev 67:197–246
72. Pancartov VA, Frenkel TM, Fainleib AM (1983) 2-Oxazolidinones. Russ Chem Rev
52:576–593
73. Kudo N, Taniguchi M, Furuta S, Endo T, Honma T (1998) Synthesis and herbicidal activities
of 4-substituted 3-aryl-5-tert-butyl-4-oxazolin-2-ones. Agric Food Chem 46:5305–5312
74. O’Hagan D, Tavasli M (1999) 1, 2-amino alcohols and their heterocyclic derivatives as chiral
auxiliaries in asymmetric synthesis. Tetrahedron Asymmetry 10:1189
75. Ager DJ, Prakash I, Schaad DR (1961) 2-Amino alcohols and their heterocyclic derivatives as
chiral auxiliaries in asymmetric synthesis. Chem Rev 96:835–875
76. Hodge CN, Lam PYS, Eyermann CJ, Jadhav PK, Fernandez Ru Y, De Lucca CH, Chang CH,
Kaltenbach RJC, Aldrich PE (1998) Calculated and experimental low-energy conformations
of cyclic urea HIV protease inhibitors. J Am Chem Soc 120:4570–4581
77. Puschin NA, Mitic RV (1937) Compounds of phosgene with hexamethylenetetramine,
m-toluidine and ethylenediamine. Justus L Ann Chem 532:300
78. Fu Y, Baba T, Ono Y (2001) Carbonylation of o-phenylenediamine and o-aminophenol with
dimethyl carbonate using lead compounds as catalysts. J Catal 197:91–97
79. Baba T, Kobayashi A, Yamauchi T, Tanaka H, Aso S, Inomata M, Kawanami Y (2002)
Catalytic methoxycarbonylation of aromatic diamines with dimethyl carbonate to their
dicarbamates using zinc acetate. Catal Lett 82:193–197
80. Bhanage BM, Fujita SI, Ikushima Y, Arai M (2004) Non-catalytic clean synthesis route using
urea to cyclic urea and cyclic urethane compounds. Green Chem 6:78–80
81. Bhanage BM, Fujita SI, Ikushima Y, Arai M (2003) Synthesis of cyclic ureas and urethanes
from alkylene diamines and amino alcohols with pressurized carbon dioxide in the absence of
catalysts. Green Chem 5:340–342
70 R.A. Watile et al.

82. Bank MR, Cadgan JIG, Thomas DE (1991) 1,3-Oxazolidin-2-ones from 1H-aziridines by a
novel stratagem which mimics the direct insertion of CO2. J Chem Soc Perkin Trans I
1:961–962
83. Xiao LF, Xu LW, Xia CG (2007) A method for the synthesis of 2-oxazolidinones and
2-imidazolidinones from five-membered cyclic carbonates and -aminoalcohols or
1,2-diamines. Green Chem 9:369–372
84. Tu M, Davis RJ (2001) Cycloaddition of CO2 to epoxides over solid base catalysts. J Catal
199:85–91
85. Sun J, Fujita SI, Bhanage BM, Arai M (2004) One-pot synthesis of styrene carbonate from
styrene in tetrabutylammonium bromide. Catal Today 93–95:383–388
86. Jagtap SR, Patil YP, Fujita SI, Arai M, Bhanage BM (2008) Heterogeneous base catalyzed
synthesis of 2-oxazolidinones/2- imidiazolidinones via transesterification of ethylene carbon-
ate with β-aminoalcohols/1,2-diamines. Appl Catal A Gen 341:133–138
87. Bigi F, Maggi R, Sartori G (2000) Selected syntheses of ureas through phosgene substitutes.
Green Chem 2:140–148
88. Getman DP, DeCrescenzo GA, Heintz RM, Reed KL, Talley JJ, Bryant ML, Clare M,
Houseman KA, Marr JJ, Mueller RA, Vazquez HML, Shieh S, Stallings WC, Stageman
RA (1993) Discovery of a novel class of potent HIV-1 protease inhibitors containing the (R)-
(hydroxyethyl)urea isostere. J Med Chem 1993(36):288–291
89. Shi F, Deng Y (2001) First gold(I) complex-catalyzed oxidative carbonylation of amines for
the syntheses of carbamates. Chem Commun 5:443–444
90. Bartolo G, Salerno G, Mancuso R, Costa M (2004) Efficient synthesis of ureas by direct
palladium- catalyzed oxidative carbonylation of amines. J Org Chem 69:4741–4750
91. McCusker JE, Main AD, Johnson KS, Grasso CA, McElwee-White L (2000) W(CO)6-
Catalyzed oxidative carbonylation of primary amines to N, N0 -disubstituted ureas in single
or biphasic solvent systems optimization and functional group compatibility studies. J Org
Chem 65:5216–5222
92. Shi F, Deng YY, Sima TL, Yang HZ (2001) A novel ZrO2–SO2– supported palladium
catalyst for syntheses of disubstituted ureas from amines by oxidative carbonylation. Tetra-
hedron Lett 42:2161–2163
93. Ion A, Parvulescu V, Jacobs P, De Vos D (2007) Synthesis of symmetrical or asymmetrical
urea compounds from CO2 via base catalysis. Green Chem 9:158–161
94. Jiang T, Ma X, Zhou Y, Liang S, Zhang J, Han B (2008) Solvent-free synthesis of substituted
ureas from CO2 and amines with a functional ionic liquid as the catalyst. Green Chem
10:465–469
95. Ogura H, Tekeda K, Tokue R, Kobayashi T (1978) A convenient direct synthesis of ureas
from carbon dioxide and amines. Synthesis 5:394–396
96. Fournier J, Bruneau C, Dixneuf PH, Lecolier S (1991) Ruthenium-catalyzed synthesis of
symmetrical N, N-dialkylureas directly from carbon dioxide and amines. J Org Chem
56:4456–4458
97. Cooper CF, Falcone SJ (1995) A simple one-pot procedure for preparing symmetrical
diarylureas from carbon dioxide and aromatic amines. Synth Commun 25:2467–2475
98. Yamazaki N, Higashi F, Iguchi T (1974) Carbonylation of amines with carbon dioxide under
atmospheric conditions. Tetrahedron Lett 13:1191–1194
99. Nomura R, Hasegawa Y, Ishimoto M, Toyosaki T, Matsuda H (1992) Carbonylation of
amines by carbon dioxide in the presence of an organoantimony catalyst. J Org Chem
57:7339–7342
100. Aresta M, Quaranta E, Tommasi I, Giannocarro P, Ciccarese A (1995) Enzymic versus
chemical carbon dioxide utilization. Part I. The role of metal centers in carboxylation
reactions. Gazz Chim Ital 125:509
101. Hayashi T, Yasuoka JEP (1998) 846679 (to Sumika Fine Chemicals Co.)
102. Fujita S, Bhanage BM, Arai M (2004) Synthesis of N, N-disubstituted urea from ethylene
carbonate and amine using CaO. Chem Lett 33:742–743
3 Indirect Utilisation of Carbon Dioxide in Organic Synthesis. . . 71

103. Fujita S, Bhanage BM, Kanamura H, Arai M (2005) Synthesis of 1,3-dialkylurea from
ethylene carbonate and amine using calcium oxide. J Mol Catal A Chem 230:43–48
104. Galli C (1992) “Cesium ion effect” and macrocyclization. A critical review. Org Prep Proced
Int 24:287–307
105. Ostrowicki A, Vogtle F (1992) In: Weber E, Vogtle F (eds) Topics in current chemistry, vol
161. Springer, Heidelberg, p 37
106. Saliu F, Rindone B (2010) Organocatalyzed synthesis of ureas from amines and ethylene
carbonate. Tetrahedron Lett 51:6301–6304
107. Jagtap SR, Patil YP, Fujita SI, Arai M, Bhanage BM (2009) Synthesis of 1,3-disubstituted
symmetrical/unsymmetrical ureas via Cs2CO3- catalyzed transamination of ethylene carbon-
ate and primary amines. Synth Commun 39:2093–2100
108. Jessop PG, Joo F, Tai CC (2004) Recent advances in the homogeneous hydrogenation of
carbon dioxide. Coord Chem Rev 248:2425
109. Federsel C, Jackstell R, Beller M (2010) State-of-the-art catalysts for hydrogenation of
carbon dioxide. Angew Chem Int Ed 49:6254–6257
110. Yu KMK, Yeung CMY, Tsang SC (2007) Carbon dioxide fixation into chemicals (methyl
formate) at high yields by surface coupling over a Pd/Cu/ZnO nanocatalyst. J Am Chem Soc
129:6360
111. Tanaka R, Yamashita M, Nozaki K (2009) Catalytic hydrogenation of carbon dioxide using Ir
(III)  pincer complexes. J Am Chem Soc 131:14168
112. Lee JS, Kim JC, Kim YG (1990) Methyl formate as a new building block in C1 chemistry.
Appl Catal 57:1–30
113. Federsel C, Jackstell R, Boddien A, Laurenczy G, Beller M (2010) Ruthenium-catalyzed
hydrogenation of bicarbonate in water. ChemSusChem 3:1048
114. Balaraman E, Gunanathan C, Zhang J, Shimon LJW, Milstein D (2011) Efficient hydroge-
nation of organic carbonates, carbamates and formates indicates alternative routes to meth-
anol based on CO2 and CO. Nat Chem 13:609
115. World Energy Council (WEC) (2010) http://worldenegy.org/wec-geis/
116. Jagadeesan D, Eswaramoorthy M, Rao CNR (2009) Investigations of the conversion of
inorganic carbonates to methane. ChemSusChem 2:878
117. Jagadeesan D, Sundarayya Y, Madras G, Rao CNR (2013) Direct conversion of calcium
carbonate to C1–C3 hydrocarbons. RSC Adv 3:7224
Chapter 4
Hydrogenation and Related Reductions
of Carbon Dioxide with Molecular Catalysts

Carolin Ziebart and Matthias Beller

Abbreviations

Bcat catecholborane
Bpin pinacolborane
bpy bipyridine
cat. catalyst
Cp* pentamethylcyclopentadienyl
Cy cyclohexyl group
DBU 1,8-diazabicycloundec-7-ene
DFT density functional theory
DHBP 4,40 -dihydroxy-2,20 -bipyridine
DHPT 4,7-dihydroxy-1,10-phenanthroline
DMFC direct methanol fuel cells
dmpe 1,3-bis(dimethylphophino)ethane
dppm 1,2-bis(diphenylphosphino)methane
dppp 1,3-bis(diphenylphosphino)propane
EMIMCl 1-ethyl-3-methylimidazolium chloride
FLP frustrated Lewis pairs
hex hexyl group
IL ionic liquid
iPr isopropyl group
IR infrared spectroscopy
KIE kinetic isotope effect
MD molecular dynamics
Me methyl group
Mes mesityl groups

C. Ziebart • M. Beller (*)


Leibniz-Institute für Katalyse Eingetragener Verein an der Universität Rostock,
Albert-Einstein Straße 29a, Rostock 18059, Germany
e-mail: Matthias.Beller@catalysis.de

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 73
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_4,
© Springer-Verlag Berlin Heidelberg 2014
74 C. Ziebart and M. Beller

MTA methanol-to-aromatics
MTD metadynamics
MTG methanol-to-gasoline
MTO methanol-to-olefins
mtppms sodium diphenylphosphinobenzene-3-sulfonate
NHC N-heterocyclic carbene
NMR nuclear magnetic resonance spectroscopy
OAc acetoxy group
OTf triflate group
Ph phenyl group
PNP pincer ligand (P donor atoms on the side arms N ~ in the middle);
analogue PNN PCP
RT room temperature
SDQ single doubles quadruples
tBu tert-butyl group
TOF turnover frequency
TON turnover number
TMP tetramethylpiperidine
TS transition state
Verkade’s base 2,8,9-triisopropyl-2,5,8,9-tetraaza-1-phosphabicyclo-
[3, 3, 3]undecane
X halide anions such as Cl Br

4.1 Introduction

In today’s chemical industry, most of the product chains are based on fossil
resources like petroleum, coal, or natural gas. Although it is known that these
resources are limited, the major part is “just” burned as fuels, and CO2 is emitted
into the atmosphere with incalculable risks for climate change [1].
Therefore, new concepts for the future have to be developed to meet the
increasing demand of energy and materials by ensuring environmental safety.
Using the abundant greenhouse gas CO2 as C1 building block to produce chemicals
and fuels would close the anthropogenic carbon cycle and save natural
resources [2].
However, the thermodynamic stability and the low reactivity of CO2 require
highly active catalysts and high energy starting materials. Nature shows that such a
transformation is practicable. By using photosynthesis, the biological reduction of
CO2 with sunlight and water to organic carbohydrates, 385∙109 t of CO2 is fixed
annually net [3]. Therefore, enzymes and different kind of mechanism have been
evolved and optimized in billions of years by nature. Biochemical studies on active
sites of these enzymes can be of benefit for understanding and developing new
synthetic catalysts capable of CO2 activation [4].
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 75

In the last decades, chemists have followed this dream and pioneering work
toward catalytic transformation of CO2 has been reported [5–7]. Especially, the
reduction of CO2 to products of higher hydrogen content is of great interest. For
example, direct hydrogenation of CO2 to MeOH with green, nonfossil-based
hydrogen is discussed to be a key step for building up a methanol economy as
MeOH can be converted into fuels as well as valuable chemicals such as olefins and
aromatics [6, 8, 9].
Despite all improvements in the last years, the catalyst costs of homogeneous
catalysts for hydrogenation of CO2 are still too high for industrial applications and
only few heterogeneous processes are established [10, 11]. However, in the last
10 years significant progress was observed that makes it more likely that industrial
processes based on homogeneous catalysis will be set up, too.
This chapter summarizes recent developments in this field starting at 2005
including hydrogen and other hydride reagents. Earlier findings have been well
presented by Jessop, Leitner, and others [12–15].
Before 2005, active catalyst systems have been developed to give HCOOH and
formamides and their mechanism have been intensively studied. However, the
separation of HCOOH from base and catalyst as well as the use of CO2 as hydrogen
storage material have been neglected—but are hot topics in this decade. Following
the principles of Green Chemistry, new systems have been developed in the last
years using green solvents, ambient conditions, bio-inspired catalysts, as well as
non-precious metals [16]. As the combustion of fossil fuels is one of the major
reasons for increasing CO2 amount in the atmosphere, the synthesis of fuels out of
CO2 using clean, nonfossil-based H2 is highly desirable [17]. Therefore, reducing
CO2 to MeOH or CH4 is of great interest, and promising catalytic systems have
been reported recently for gaining MeOH. Homogeneous catalysts capable of
hydrogenating CO2 to CH4 with H2 are still unreported, but interesting activities
have been achieved using a combination of boranes and silanes.

4.2 Hydrogenation of CO2/HCO3 to Formic Acid


Derivatives

HCO3  þ H2 ! HCO2  þ H2 O
ROH
CO2 þ H2 !base
HCO2 H  base ! HCOOR
H2 O ð4:1Þ
HNR2
CO2 þ H2 ! HCONR2
H2 O

The direct hydrogenation of CO2 to HCOOH is an endergonic reaction. Therefore,


base is often applied to drive the thermodynamics by proton transfer. Available
products are formates, formic acid–base adducts, as well as alkyl formates in the
presence of alcohols and formamides by using primary or secondary amines (Eq. 4.1).
76 C. Ziebart and M. Beller

Catalytic systems reported before 2005 are usually noble-metal systems with mono- or
bisphosphine ligands as well as a few examples of bipyridine-based ligand structures.
The best catalyst presented until 2005 is the (PMe3)4RuCl(OAc) catalyst from Jessop
which hydrogenates CO2 at 50  C and 200 bar of H2/CO2 with high TON of 32,000
and an astonishing TOF of 95,000 h1 [18].
Since then, new ligand structures have been developed and applied successfully
in the hydrogenation of CO2 to formic acid derivatives. This section gives an
overview of recent results subdivided by ligand classes.

4.2.1 Using Mono- and Bidentate Phosphine Ligands

As mentioned above, the Ru trimethylphosphine catalyst system [18] was one of the
most active ones until 2005. Regarding the high interest in understanding the
mechanism, several computational and analytical studies have been reported iden-
tifying possible reaction species (Scheme 4.1). All of them have in common that a
Ru dihydride species 1 with four phosphine ligands exists. The dissociation of one
phosphine ligand leads to cis-Ru(H)2(PMe3)3. Here, Sakaki started his investiga-
tions with DFT and MP4(SDQ) methods to figure out differences between this real
catalyst and the model catalyst cis-Ru(H)2(PH3)3 [19]. As the PMe3 group is more
donating than its analog, the trialkyl phosphine ligand showed a stronger trans-
influence than PH3. Both, strong donating ligands and polar solvents, made the CO2
insertion into the metal-H bond more favorable. This step is proposed to be the
rate-determining step in the hydrogenation of CO2 with 2 in absence of water. The
catalytic cycle can be described in the following steps: CO2 insertion into Ru(II)-H,
isomerization of the ruthenium(II)-η1-formate intermediate 3, six-centered σ-bond-
metathesis of the η1-formate with a dihydrogen molecule 4, and dissociation of
formic acid to give again the dihydride species (Scheme 4.1, phosphine dissociation
mechanism (a), phosphine groups are drawn as P for clarity).
Taking into account that water was observed to accelerate the catalytic system,
the group of Sakaki presented a second computational study using cis-Ru
(H)2(PMe3)3(H2O)2 as catalytic intermediate [20]. They figured out that aqua
ligands can accelerate the nucleophilic attack of the hydride ligand on the CO2
molecule by forming hydrogen bonds (6). The resulting ruthenium(II)-η1-formate
3 isomerizes fast in the presence of water which suppresses the deinsertion of CO2.
The rate-determining step now is believed to be the coordination of the H2 molecule to
the ruthenium(II)-η2-formate complex. The last step, the heterolytic H2 cleavage, is
facilitated again by water molecules (7). However, Jessop’s NMR results have been
inconsistent with such a phosphine dissociation mechanism. For this reason, he
proposed a different cycle based on the unsaturated, cationic ruthenium species
[(PMe3)4RuH]+ 8 [21]. This mechanism also explains the necessity of the added
base. The base on the one hand traps the formic acid and on the other hand helps to
form the active catalytic species by abstracting a hydride ligand leading from 1 to 8.
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 77

a
-HCOOH, + CO2
H H
O
O P O OH Phophine dissociation
P -P P
C PH P mechanism
P H + CO2 O O H O H H
Ru H Ru O
Ru Ru
P H Ru P H P H P H
P H without water
P 1 2 P P 4 P 5
P 3
b +H2O
-P H
+ H2O O H with water
O
+ CO2
H O
P O H
H C P
P O H O H
O H O
Ru -H2O Ru Ru H
P H P H P H
P 6 P 3 P 7

O
+ HbaseX P X P X P X P X
C H base-assisted
- H2, - base P + CO2 P O P +H2 P
Ru Ru H formation of
P H P H Ru Ru
P O2CH P O2CH cationic Ru-species
P 8 P 9 P 10 P 11
-HCOOH

Scheme 4.1 Two different reaction pathways for the (PMe3)4RuCl(OAc) system proposed by
Sakaki and Jessop

Most recently, our group synthesized benzoyl- and naphthoyl-substituted


phosphines which are stable to air and moisture [22]. The related Ru complexes
showed good activity in the bicarbonate reduction with H2 at 80  C (TON up to
1,600). CO2 as well as carbonyl compounds could be hydrogenated, too.
Using simple phosphite ligands at the Ru center such as [trans-RuCl2{P(OMe)3}4],
the group of Thiel achieved good TON up to 6,630 and TOF of 1,655 h1 under
supercritical conditions [23]. This was the first time phosphite ligands have been
shown to form active complexes for the hydrogenation of CO2. However, the appli-
cation of the expensive base DBU in the presence of C6F5OH is a drawback and an
exchange through cheaper bases is highly desirable.
Taking advantage of the water-accelerating effect, Zhao and Joó published a
water-soluble [RhCl(mtppms)3] catalyst capable of CO2 hydrogenation to free
formic acid in aqueous HCOONa solutions [24]. However, only low TON < 150
was observed after 20 h at 100 bar pressure (CO2/H2 – 1:1) and 50  C. This might be
due to the low basicity of the formate.
The hydrogenation of CO2 to HCOOH is reversible. Hence, many catalyst systems
have been developed which efficiently and selectively decompose HCOOH to hydro-
gen and CO2 [25–27]. Hydrogen, ideally produced by renewable energy sources, for
example, by electrolysis of water, is a potentially explosive, volatile gas which is
usually stored in high pressure tanks [28–31]. By reacting it with CO2 or bicarbonate,
H2 can be stored and transported easily as liquid or solid and H2 can be released easily
if needed. Therefore, researchers developed different systems capable of storing
hydrogen based on CO2/HCOOH or bicarbonate/formate. Both the groups of Papp
and Joó [32] as well as our group [33] presented at the same time catalyst systems
able of recharging H2 based on a bicarbonate/formate cycle. In comparison to CO2,
bicarbonate is well soluble in water and its solutions are easy to handle and can be
converted catalytically under mild conditions. As these systems work amine-free, no
base evaporation is possible with the H2 stream.
78 C. Ziebart and M. Beller

Hydrogen release

HCOOH NEt3 CO2 in


in DMF [RuCl2(benzene)]2 NEt3/DMF
dppm HCO3- in
HCO2- in
water water

Hydrogen storage

Scheme 4.2 One catalyst—two hydrogen storage systems based on CO2 or bicarbonate by the
group of Beller

Papp and Joó observed reversibility with the [RuCl2(mtppms)2]2 complex in


aqueous bicarbonate solution in a sapphire NMR tube [32]. They showed three
cycles of both the hydrogenation of bicarbonate to formate and the dehydrogenation
of formate to bicarbonate. However, the decomposition of the formate slows down
significantly at around 50 % conversion. Therefore, only half of the formate amount
can be used for hydrogen storage in this system. Higher yields for the dehydroge-
nation of formates at even lower temperature (70 % at 40  C) were shown by our
group with a [RuCl2(benzene)]2 precursor in the presence of 4 eq of dppm as ligand
[33]. In comparison to earlier studies [34], the yields for the bicarbonate reduction
could be improved to 96 % at 70  C, using 80 bar of H2 in a water-THF mixture
without additional CO2. A good TON of 1,108 was achieved after 2 h. Exchanging
the sodium cation through NH4+ or Li+, K+, Mg2+, and Ca2+cations, the productivity
decreases and often a significant amount of CO2 is released with the H2 during the
formate decomposition. Our group also demonstrated that the catalyst is capable of
hydrogen storage at RT based on CO2/HCOOH∙NEt3 (Scheme 4.2). The catalyst
system runs eight cycles without showing significant decrease in activity [35].
For the hydrogenation at RT, high formic acid amine ratios up to 2.31 were
observed with TON up to 3,200. In addition, we could present an astonishing TON
up to 800,000 for continuous H2 liberation using [RuCl2(benzene)]2 and dppe.
However, it should be noted that the gravimetric hydrogen content of formic acid
is limited to 4.4 % which is further reduced by the addition of amine or solvent.
To clarify possible reaction mechanisms for complexes with bidentate phosphine
ligands, ab initio MTD and MD studies were performed by the group of Baiker for the
dihydride catalyst [Ru(dmpe)2(H)2] 13 [36]. This study was supported by further
DFT calculations as well as IR and NMR experiments [37]. The authors concluded
the coordination of molecular H2 to the complex is the rate-determining step. That
requires dissociation of the formate from the Ru center 14 to leave a free coordination
site for the H2. The formate anion then interacts weakly with the Ru(H2)+ complex 15
(Scheme 4.3). The activation energy for the H2 adduct was found to be lower for the
complex where the two hydrides are in trans position to each other (13). However, in
solution more than 95 % of the cis-complex 12 was observed. The CO2 insertion is
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 79

H
- HCOOH
O O O
H
H
P H O H
trans + CO2 + H2
P H P P P P P P
Ru
P
Ru H
cis P Ru P P Ru P P P
P 12 H 13 H H
14 15

Scheme 4.3 Proposed mechanism for the [Ru(dmpe)2(H)2] catalyzed CO2 hydrogenation by
Baiker

believed to happen via a concerted mechanism where the Ru-H bond breaks at the
same time as the Ru-O-bond is formed. This can be imagined as a “formate ion”
rotation.
Based on kinetic measurements by using stop-flow technique, Schindler’s group
calculated activation parameters for a Rh catalyst involving [(dppp)2RhH] as the
active species [38]. For this complex, the following mechanism was proposed: the
formation of an ionic Rh(I) formate complex through insertion of CO2 into the Rh-H
bond and subsequent dissociation of formate, followed by oxidative addition of H2 to
give an ionic Rh(III) dihydride formate complex which under reductive elimination
of formic acid yields again the neutral Rh(I) complex.
In addition to catalyst variations and mechanistic studies, different kinds of
concepts have been developed to solve the problem of the formic acid separation
from base and catalyst. For example, the groups of Han and Leitner presented
independently two systems based on ionic liquids. That the equilibrium of CO2/H2
to HCOOH is more favored in IL than in H2O, due to the strong solvation of
HCOOH in IL, was shown by Nakahara [39]. The idea of Han was to use an IL with
an extra amine function 16 in combination with a ruthenium catalyst bound on silica
[40]. By inserting a second amine function on the IL (17), the activity could be
increased from 100 h1 up to 920 h1 TOF with a TON of 1,840 using 90 bar of H2
and a total pressure of 180 bar [41]. Both systems were recycled four times using
the procedure described in Fig. 4.1. Addition of water accelerates the reaction,
maybe due to a decreased viscosity, interaction of water in the catalytic cycle, or
bicarbonate formation as the true substrate.
Leitner and co-workers were using the IL as stationary phase containing the
precursors [Ru(cod)(methallyl)2] and PBu4+mtppms as well as additional base
[42]. The formic acid extraction occurs with supercritical CO2 in a continuous-flow
single process unit where scCO2 serves as mobile and extracting phase at the same
time. With the base NEt3 and EMIM Cl as IL, a TON of 516 was achieved after half
an hour at 50  C and 50 bar of both CO2 and H2. By exchanging the anion of the IL
through HCO2 to give 18, increased TON and TOF were observed.
Another interesting concept for the production and separation of pure formic
acid was presented by Schaub and Paciello [43]. Key to success is the skillful
choice of solvents by taking advantage of the diverse solubility and miscibility of
all components (Fig. 4.2).
80 C. Ziebart and M. Beller

130 °C
H2 CO2 HCOOH 110 °C N2 flow
Filtration HCOOH HCOOH
IL H2O IL H2O IL [41, 42]
- cat. IL H2O IL
cat.(s) cat.(s) -H2O
HCOOH

or HCOOH extraction with scCO2 as mobile phase and IL as stationary phase [43]

F3CSO3- F3CSO3- HCO2-


N N +
N +
N N N+
N N N

16 17 18

Fig. 4.1 Concepts for HCOOH separation using IL presented by Han and Leitner

1. Hydrogenation 2. Catalyst extraction 3. Product separation


(by destillation)

CO2
+ H2 HCOOH

NHex3 amine
NHex3 + + cat. recycling NHex3
Ru-PnBu3-cat. Diol +
Diol + cat. traces HCOOH NHex3
Diol
HCOOH NHex3

Diol recycling

Fig. 4.2 Liquid-liquid phase process for formic acid formation and separation by Schaub and
Paciello [43]

Here, in the first step, CO2 is hydrogenated with the [Ru(H)2(PnBu3)4] catalyst in
the presence of NHex3 and a diol. The formed formic acid salts [NHex3∙HCOOH]
are not miscible in free amine, but good soluble in the diol phase. At the same time,
most of the lipophilic catalyst stays in the unpolar amine phase. To get rid of the
remaining catalyst traces, the diol phase is extracted with the resulting amine of the
formic acid salt cleavage. The amine is then recycled to the hydrogenation vessel,
whereas the formic acid salt enriched diol phase is treated thermally under mild
conditions to give pure formic acid and free NHex3 which forms again a two-phase
system with the diol. Both phases are back transferred to close the process cycle.

4.2.2 Using Bipyridine-Based and Carbene-Type Ligands

A second ligand class known for CO2 hydrogenation is based on the bipyridine
structure. In this context, the groups of Ogo and Fukuzumi started kinetic
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 81

O OH

N N
2H+
M M
-2H+
N N

O OH
Pyridinolate form Pyridinol form

HCO3- to HCO2- no hydrogenation


pH>8 pH<7
homogeneous heterogeneous

Scheme 4.4 pH switchable catalysts for CO2 hydrogenation reported by Himeda

investigations of Ir(III) and Ru(II) bpy complexes in citric buffer solutions


(pH ¼ 3) to obtain pure formic acid [44]. By comparing TON in dependence of
H2 or CO2 pressure, they concluded that in the case of ruthenium, the H2 coordi-
nation to the corresponding aqua complex is rate determining. This is in contrast to
the iridium complexes. Here, the authors believe that the rate-determining step is
the CO2 insertion into the active hydride complex.
By inserting hydroxy groups in para- or ortho-position at the pyridine rings,
interesting behavior of the catalysts in dependence of the pH was reported. This
ligand class shows an acid–base equilibrium between the pyridinol and pyridinolate
form (Scheme 4.4) which influences the catalytic activity and water solubility.
In 2007, Himeda presented different catalysts for the CO2 hydrogenation in
aqueous KOH solutions to potassium formate using Ru, Ir, and Rh half-sandwich
complexes bearing DHPT or DHBP ligands [45]. Under basic conditions, the
deprotonated, water-soluble form exists showing strong electronic donation prop-
erties. High TON up to 220,000 was observed for an Ir DHPT complex at 120  C
and 60 bar of each CO2 and H2 in 48 h. Lowering the pH to acidic conditions, the
pyridinol form of the catalyst predominates. The catalyst becomes insoluble in
water and precipitates. Now, the catalyst can be reused easily and showed
maintaining activity even after four cycles. To confirm the importance of the
formation of the pyridinolate species, the hydroxy groups have been methylated
resulting in significantly less active complexes. Other groups in para-position like
carboxy or methyl groups are inferior, too [46]. For further details see review [47].
To study the effects of the ligand structure in more detail, Himeda and Fujita
synthesized new iridium half-sandwich complexes differing in position and quan-
tity of the hydroxy groups (Scheme 4.5).
82 C. Ziebart and M. Beller

+2 +2 +2 +2
OH HO HO OH HO OH

N N N N H 2O N N
Cp* Cp* Cp* Cp*
Ir < Ir < Ir Ir = Ir
H2O H 2O H 2O Cp* H2O
N N N N N N

OH HO HO OH HO OH
19 20 21 22

Scheme 4.5 Ligand structure study on the reactivity of Ir half-sandwich complexes by Himeda

Moving the hydroxy groups from para to ortho position (19–20) resulted in
a rate enhancement by a factor of 2 comparing the initial TOF after 10 min
[48]. Doubling both the quantity of the hydroxy groups as well as the amount of
Ir centers 21, another increase in the reactivity in order of 2.5 was observed for the
hydrogenation of CO2. Surprisingly, the dinuclear complex 22 showed similar
activity as the mononuclear one with four hydroxy groups 21. To explain these
results, DFT studies and deuteration experiments have been performed [49]. The
computational study identified the water-assisted heterolysis of H2 to be the rate-
determining step. This is in line with the NMR experiments as kinetic isotope
effects of H2O, H2, and bicarbonate have only been observed for the ligands
containing hydroxy groups in ortho position, not para. This is the first time a
kinetic isotope effect was proved for the hydrogenation of CO2, revealing that
H2O is involved in the rate-determining step. The significant increase in reactivity
is believed to be caused by the high σ-donor strengths of the four O groups and the
accelerated proton transfer by forming water bridges.
Using the binuclear Ir complex 21, Himeda, Fujita, and Hull showed that their
system is not only capable of CO2 hydrogenation but also can decompose HCOOH
to CO2 and H2 under acidic conditions [50]. High TON of 308,000 and very good
TOF of 228,000 h1 were achieved for the decomposition at 80–90  C in the
presence of HCOOH/HCOONa (1:1). Increasing the pH to basic conditions, the
same catalyst hydrogenates CO2 with TON of 153,000 and TOF of 53,800 h1 in
the presence of bicarbonate at 50  C/80  C. To rule out bicarbonate as only
substrate, a test reaction without CO2 was run yielding only a low formate produc-
tion. However, long reaction times and the necessity of adding stoichiometric
amounts of base or acid to switch the reaction make it so far unpractical for
industrial application.
The same concept of hydrogen storage was reported by Fukuzumi using the
water-soluble, proton-switchable phenylpyrazolyl organoiridium aqua complex 23
(Scheme 4.6) [51]. Both the hydrogenation of CO2 under slightly basic conditions
at RT under atmospheric pressure of H2 and CO2 and the formic acid decomposition
under acidic conditions are possible. Due to KIE effects, the authors concluded that
the rate-determining step is the formic acid decomposition via β-hydride elimina-
tion of the formate to release CO2. In case of the hydrogenation of bicarbonate, both
the catalyst and bicarbonate are believed to be involved in the rate-determining step
as the TOF raises with increasing amount of bicarbonate. A more general view on
these Ir-H complexes is presented in a review [52].
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 83

PMe3

N N N Ru
Ir OH2
O Cl
N O Ru
N Cl PMe3
OH
PMe3
23 24 25

SO3-K+ SO3-K+
N N
I I
N O N O
N
Ru Ir Ir
Cl N O N O
I I
N
N N N
N

26 27 SO3-K+ 28 SO3-K+

Scheme 4.6 Catalysts capable of HCOOH production by Fukuzumi, Thiel, Süss-Fink, and Peris

Further systems capable of CO2 hydrogenation were presented by Niedner-


Schatteburg, van Wüllen, and Thiel [53] using catalyst 24 and the group of Süss-
Fink [54] including the arene ruthenium oxinato complex 25. However, in all of
these cases only moderate TONs were achieved.
Another interesting ligand class for the activation of CO2 was discovered by
Peris. A series of bis-N-heterocyclic carbene complexes with Ru 26 and Ir 27 were
tested in both CO2 hydrogenation and transfer hydrogenation [55]. As these
chelating NHC ligands are known for their high thermal stability, reaction temper-
atures up to 200  C are tolerated. Interestingly, if the carbene position is blocked by
a methyl group, a bis-abnormal coordination of the NHC ligand to the iridium
center is observed [56]. This complex 28 is even more active and the highest known
TON of 1,320 was reported for transfer hydrogenation of CO2 using iPrOH in
aqueous KOH solutions.

4.2.3 Using Pincer Ligands

For more than 30 years, pincer-type ligands have been known [57]. The high
stability of this complexes are due to the strong coordination between the tridentate
ligand and the metal center. At the same time, they offer interesting redox chemistry
as these noninnocent ligands are directly involved in the reaction mechanism. This
enabled new and unknown reactivities. Catalytic applications of these ligand-metal
cooperations are, for example, arylation and coupling reactions, alkane dehydro-
genations, hydroaminations, and a bunch of hydrogenation reactions [58–61].
84 C. Ziebart and M. Beller

O
C O
H H H H
i Pr2P Ir Pi Pr2 i Pr2P Ir Pi Pr2
H N +CO2 H N

29 30 O
C H
H H H
-H2O O
i Pr2P Ir Pi Pr2 i Pr2P Ir Pi Pr2
H N H H N
H H OH-
i Pr2P Ir Pi Pr2
36 31
H N H
-HCOO-
37
H H H
+H2
iPr2P Ir Pi Pr2 + OH- i Pr2P Ir Pi Pr2
N H N

35 32

+OH-
-H2O
H H
OH2 OH
i Pr2P Ir Pi Pr2 i Pr2P Ir Pi Pr2
H N H N

34 33

Scheme 4.7 Possible reaction mechanism for the Ir-trihydride-pincer complex of Nozaki

Recently, the first example of CO2 hydrogenation to formate with this ligand class
was presented by Nozaki. This system is highly efficient and was highlighted in 2010
[62]. In fact, Nozaki’s Ir(III)(H)3-pincer complex 29 achieved astonishing TON up to
3,500,000 in 48 h at 120  C and TOF of 150,000 h1 after 2 h at 200  C using each
30 bar of CO2 and H2 in an aqueous KOH solution of 1 M [63]. The proposed
mechanism is presented in Scheme 4.7. The first step is the CO2 insertion into Ir-H
bond of 29 forming complex 30 with the formato group in cis position to the N atom.
After dissociation of the formate, the Ar-CH2-PiPr2 moiety of the ligand is
deprotonated followed by dearomatization of the pyridine ring. To this amido iridium
dihydride complex 35, a hydrogen molecule coordinates which is then heterolytically
cleaved to give the full-aromatized Ir(H)3 29 back. Yang reported a different
pathway by studying computationally PNP-pincer complexes with Ir(H)3 and Fe/Co
(H)2(CO) [64]. He claimed that the OH-triggered H2 cleavage (37) is more favorable
than Nozaki’s aromatization/dearomatization mechanism as well as Ahlquist’s sug-
gestions based on DFT studies [65]. In response Nozaki and Morokuma presented
their own computational study for the CO2 hydrogenation with complex 29 [66].
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 85

As a result, both pathways, the aromatization/dearomatization reaction as well as the


hydrogenolysis suggested by Yang, are possible as their transition state energies are
very similar. Deeper insight in the mechanism was provided by Hazari who examined
the trans-influence of various ligands for insertion of CO2 into the Ir-H bond of 29
[67]. The strongest trans-effect was observed by the hydride ligand which is believed
to weaken the Ir-H bond by increasing the nucleophilicity of this hydride making the
CO2 insertion more favorable.
Other non-considered mechanisms were pointed out by Milstein’s group. They
proposed a double deprotonation by methylating a Ni(II) PNP complex where the
negative charge is delocalized in the amido-type backbone [68] as well as a
methylated Co(I)-pincer complex which lost a H radical by homolytic C-H bond
cleavage [69]. Here, the unpaired electron is delocalized in the ligand backbone and
evidence for the remaining aromaticity of the pyridine ring was detected. The
abstraction of the H radical is believed to be caused by the solvent as an increase
of the reaction rate was observed in aliphatic solvents.

4.2.4 Using Non-precious Metals

Despite increasing scientific interest in the hydrogenation of CO2, the use of


bio-relevant metals such as iron, manganese, or cobalt as catalysts has scarcely
been investigated compared to noble-metal-based systems. Before 2005, only a few
examples were reported yielding higher TON than 10 [70]. However, long reaction
times and harsh conditions are needed.
In 2011, Yang predicted computationally an iron-pincer complex to be capable of
CO2 hydrogenation [64]. In the same year, Milstein presented an analog Fe complex
38 (Scheme 4.8) not only active in CO2 hydrogenation but also in the bicarbonate
reduction to formate [71]. In aqueous solutions, bicarbonate can be converted to the
corresponding formate at ambient pressure and 80  C with a moderate yield 30 % and
a TON of 320. Using CO2, even higher TON up to 600 was achieved. Based on IR
and NMR experiments, a similar mechanism as for the Ir(H)3-pincer complex 29 is
proposed.
Whereas mono-, bi-, and tridentate phosphine ligands are well investigated for
noble metals in the catalytic hydrogenation of CO2, tetradentate ligands have not
been presented at all. The application of the tetradentate phosphine ligand
P(CH2CH2PPh2)3 in combination with Fe(BF4)2∙6H2O in MeOH led to the highly
active complex 39 [72]. With this complex, our group showed that bicarbonate can
be reduced to formate in high yields (88 %) and TON (up to 610) at 80  C and
60 bar of H2. Testing a variety of bidentate and tridentate ligands which failed all
in generating an active catalyst in situ, it seemed to be that iron needs a very
well-defined environment. Even other iron sources instead of Fe(BF4)2∙6H2O gave
significant lower yield. Replacing bicarbonate by CO2, methyl formate was produced
in the presence of NEt3 and the corresponding formamides were observed by using
dialkyl amines as base yielding TON up to 730.
86 C. Ziebart and M. Beller

Pt Bu2
BF4-
H P Me2 H Me2
Ph2P M PPh2 P P
N Fe CO P
Co
Ph2P Ph2P Fe PPh2 P P
H H Me2 Me2
M = Fe, Co Ph2P
Pt Bu2 F BF4-
38 39 40 41

Scheme 4.8 Non-precious metals active in CO2 hydrogenation published by Milstein, Beller, and
Linehan

Changing the metal from Fe to Co, slightly higher yields were observed for the
bicarbonate reduction and higher temperature was tolerated. In addition, lower H2
pressure is applicable [73]. TON up to 3,900 for sodium formate and 1,300 for DMF
could be detected. By synthesizing Co monohydride and Co dihydrogen complexes
and testing them in the catalytic hydrogenation of bicarbonate, the monohydride
complex was excluded as possible active species, whereas the Co dihydrogen com-
plex was active. These results were in correlation with the performed NMR studies.
By synthesizing the phenyl-bridged analog of the tetradentate phosphine ligand
mentioned above and combining it with Fe(BF4)2∙6H2O, the most active and stable
iron catalyst 40 known for the hydrogenation of bicarbonate and CO2 was developed
[74]. The crystal structure of 40 exhibited an unusual Fe-F bond, where the F
originally came from the BF4 anion. Both the in situ system and the well-defined
complex 40 showed the same activity, resulting in TON up to one order of magnitude
higher than the previously reported Fe systems for the bicarbonate reduction (TON
7,550). With higher catalyst loading, the reaction reached its equilibrium already after
5 h with TOF up to 770 h1. Furthermore, the hydrogenation of CO2 in the presence
of base gave formic acid, methyl formate, and formamides, the latter in TON up to
5,100. NMR studies indicated that a Bianchini-type complex [75] has been formed
being in equilibrium with a Fe-hydride complex where the dihydrogen ligand has
been exchanged by a solvent or a CO2 molecule. Through base-assisted heterolytic
hydrogen cleavage, an active dihydride Fe complex was proposed to be formed into
which CO2 inserts to give the hydride formate complex. Dissociation of the formate
followed by the coordination of a dihydrogen molecule led back to the Bianchini-type
complex and HCOOH∙base adduct.
Very recently, the group of Linehan published NMR experiments where Co(I)
(dmpe)2H shows very high TOF up to 74,000 h1 for the hydrogenation of CO2 to
formate [76]. However, by using other bases than the expensive and very unusual
Verkade’s base, the activity dropped down dramatically. Even with much higher
catalyst loading and the use of the expensive, in CO2 hydrogenation successfully
tested base DBU, almost no activity was observed. A scale-up of this process is
therefore very unlikely.
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 87

Table 4.1 Reported catalysts for the hydrosilylation of CO2 to silyl formate
Entry Catalyst T [ C] CO2 [atm] TON TOF [h1] Ref.
1 Ru2Cl5(MeCN)7 40 14 4,619 231 [77]
2 cis-[RuCl2(MeCN)4] 85 10 g – 3,700 [78]
3 Iridium complex RT 3 112 0.8 [80]
4 Zinc complex 100 6.8 1,006 2.9 [81]
5 Cu bisphosphine complex (Ph) 60 1 8,100 1,350 [82]
6 [(NHC)Cu(OtBu)] 60 1 7,489 1,248 [83]
7 Cu bisphosphine complex (iPr) 60 1 70,000 2,917 [79]
8 Rh2(OAc)4 50 1 180 90 [84]
9 [Et3SiPh][B(C6F5)4] 30 1 – – [85]

4.2.5 Using Other Hydride Reagents

Parallel to the hydrogenation of CO2, catalysts for the reduction of CO2 to silyl
formate with Hsilanes have been reported. The most recent systems are listed in
Table 4.1.
In case of the Ru2Cl5(MeCN)7, the catalyst could be recycled 10 times, resulting
in an overall TON up to 4,600 [77]. The same group achieved with the similar
catalyst cis-[RuCl2(MeCN)4] and MePhSiH high TOF of 3,400 h1 yielding 96 %
of silyl formate at 85  C [78]. The highest TONs were achieved by the copper
(1,2-bis(diisopropylphosphino)benzene) complex [79] with a very good TON of
70,000 after 24 h at 1 atm of CO2 pressure.
From a mechanistic point of view, two different pathways are proposed
depending on how CO2 inserts. For the systems [79] and [80], in the first step, the
silane is activated on the metal center resulting in the formation of a metal-hydride
complex and the transfer of the silyl cation to an anionic ligand (Cl or triflate). In
the second step, the silyl cation activates CO2 and subsequent silyl formate is
reductively eliminated (outer-sphere mechanism).
The other systems [79, 81–83] are believed to follow a different mechanism. Here,
the active catalytic species is formed by the reaction of the precursor and silane to yield
the corresponding metal-hydride complex. Into this metal-H bond inserts a CO2
molecule to build up the metal formate complex. In the last step, this formate complex
reacts with silane to regenerate the metal-H species and release silyl formate.
By adding water to silyl formate, formic acid is formed. This transformation is not
really competitive with the reduction of CO2 to formic acid using molecular hydrogen
as for the latter even more active catalysts systems are known. However, using other
nucleophiles, such as amines or Grignard reagents, offers new synthetic ways to a
number of synthesis products (Scheme 4.9). For example, benzhydrol was formed in
77 % yield by reacting silyl formate with two equivalents of PhMgBr [84].
A stoichiometric metal-free reduction of CO2 with trialkyl silanes and stoichio-
metric amounts of trityl borate was shown by the group of Müller [85]. They
observed different products depending on the solvent applied. In PhCl, disilylated
formic acid or disilyl methyl oxonium is formed. By quenching these intermediates
88 C. Ziebart and M. Beller

R2 N H
R2NH

cat. O O OH
ArMgBr ArMgBr
R3Si-H + CO2 R3Si
O H Ar H Ar Ar

H2O O

HO H

Scheme 4.9 Reduction of CO2 to silyl formate and possible applications presented by Mizuno

with water, HCOOH and MeOH are obtained. In contrast, in PhH, the transformation
of benzylic cations gives silyl ester or benzoic acid.
An example for a hydroboration of CO2 to formate was presented by Labinger
and Bercaw [86] using trialkyl boranes in the presence of Rh and Ni dmpe
complexes to form formate-borane adducts. However, closing the catalytic cycle
through breaking the formate-borane adducts failed.
In 2013, the group of Shintani and Nozaki presented a copper-NHC-catalyzed
hydroboration of CO2 to formic acid by quenching the reaction mixture with HCl
[87]. This way, yields up to 87 % of formic acid have been observed. By adding
primary or secondary amines to the reaction mixture instead of quenching with
HCl, they isolated the corresponding formamides in good to very good yields
(81–98 %). This N-formylation is also possible in one step by reacting CO2 with
pinacolborane in the presence of amine. However, the yields decreased slightly
(e.g., from 84 to 71 % for p-anisidine).

4.3 Hydrogenation of CO2 to MeOH

Searching for alternative sources of fuels and energy storage mediums, MeOH is
considered for several good reasons. First, it can be dehydrogenated to give H2 and,
therefore, can serve as hydrogen storage carrier [88, 89]. Second, methanol has a
high energy content of 22.7 MJ kg1, making it suitable for energy storage
(CH4: 24.3 MJ kg1) [6]. It can be used as liquid fuel (blended with gasoline or
directly in direct methanol fuel cells, DMFC [90]) or can be converted to gasoline
(MTG process). Last but not least, MeOH is a valuable feedstock as it can be
transformed into ethylene or propylene in the MTO (methanol-to-olefins) process
or to aromatics in the MTA (methanol-to-aromatics) process using zeolites. There-
fore, MeOH covers all important basic chemicals for a wide range of product chains
[6, 91]. The advantage of such a MeOH economy in comparison to the often
discussed H2-based technology is that MeOH as a liquid can be used analogous
to petroleum and therefore can be transported and stored easily in the maintaining
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 89

infrastructure. In addition, MeOH is highly H2 enriched (12.6 w%) in comparison to


many other H2 storage devices.
The most sustainable way for producing MeOH out of CO2 and water combines
the electrolysis of water to hydrogen and the subsequent hydrogenation of CO2 to
MeOH. The cost of this process will strongly depend on the electricity price for
supplying the needed hydrogen amount. However, even today, this process can be
competitive regarding the highly increased prices of gasoline in European countries
[8]. The first commercial CO2-to-renewable-methanol plant based on geothermal
sources was established in 2012 in Iceland, producing annually 3,500 t of MeOH,
and is planned to be expanded [9]. Industrial applications of homogeneous catalyst
systems in this field are still far away from reality, but the first recent examples of
the homogeneous catalyzed hydrogenation of carbon dioxide in the last 3 years are
promising.

4.3.1 Catalytic Reduction with Hydrogen

CO2 þ 3 H2 ! CH3 OH þ H2 O ð4:2Þ

One of the first concepts for the hydrogenation of CO2 to MeOH (Eq. 4.2) was
presented by the group of Sanford [92]. They set up a cascade by using three
different catalysts: Ru(PMe3)4Cl(OAc) for the hydrogenation of CO2 to formic
acid, Sc(OTf)3 as catalyst for the esterification of the formic acid to methyl formate,
and finally the Ru PNN-pincer complex 42 which converted methyl formate to
methanol. Key to success was the separation of the last catalyst from the second one
to avoid deactivation. Therefore, methyl formate is transferred by a temperature
ramp to an outer vessel containing the pincer complex 42. The overall TON of
21 for the transformation of 13CO2 to MeOH is low, but showed that this reaction is
achievable by homogeneous catalysis. As one reason for this low activity, Sanford
recognized an inhibition of the Ru PNN-pincer complex 42 by CO2 [93]. Two
products were observed under CO2 atmosphere (Scheme 4.10). The kinetic product
43 is observed immediately under 1 atm of CO2 by forming a C-C bond between
the CO2 and the phosphine arm. This process is reversible. By simply keeping the
complex longer under CO2 atmosphere or heating the solution, the irreversible
product 44 is formed at the amine arm of the pincer complex and single crystals
could be grown from the solution.
Milstein observed a similar phenomenon with a Ru PNP ligand with two
phosphine arms [94]. Here, only a reversible addition of CO2 to the phosphine
arm was found confirming Sanford’s results. Calculations reveal no solvent effect
on the TS energies, indicating a concerted mechanism and the absence of a charge
separation like Ru+ COO. Another example of CO2-ligand interactions was
shown by Oro and Langer [95] using an Ir(I) dppm catalyst. Here, CO2 inserts
90 C. Ziebart and M. Beller

H H H

Pt Bu2 Pt Bu2 Pt Bu2


N 1 atm CO2 N N
Ru Ru Ru
O
N CO N N
Et2 Et2 CO Et2 CO
42 O 43 O 44
kinetic product thermodynamic
O product

Scheme 4.10 CO2 interactions with the Ru PNN-pincer complex observed by Sanford

PPh2
Ph2P
Ru
P HNTf2
45 Ph2
3 H2 + CO2 MeOH + H2O
60 atm 20 atm 140 °C, 24 h, TON up to 221

Scheme 4.11 The first homogeneous one-catalyst system for the reaction of CO2 to MeOH
reported by Leitner

into the C-H bond of the bisphosphine ligand to give the binuclear complex [IrCl
(dppm)(H){(Ph2P)2C-COOH}]2.
The first one-catalyst system for the production of methanol was reported by
Klankermayer and Leitner [96]. Using the Ru catalyst 45 in the presence of 1 eq
HNTf2, MeOH is produced with a TON of 221 in a THF/EtOH mixture at 20 bar of
CO2 and 60 bar of H2 (Scheme 4.11). They propose a cationic Ru hydrido
dihydrogen complex to be the active species.
In addition to the systems mentioned above, a few concepts for the indirect
conversion of CO2 to MeOH were developed, reducing substrates which can be
produced from CO2 (Scheme 4.12). Pioneering work has been published by
Milstein. Using the Ru-pincer complexes 42 or 46, alkyl formates, organic carbon-
ates, and carbamates [97] as well as challenging ureas [98] were converted under
mild conditions in quantitative yield, often for the first time. Key to success is the
unique behavior of the pincer complex being involved in the mechanism by metal-
ligand cooperation. This work was also highlighted by Dixneuf [99] and Choudhury
[100]. Notably, the activities need to be improved for industrial application and the
economy of the process will also depend on H2 cost.
One year later, Yang presented a DFT study on the dimethyl carbonate reduction
to MeOH with the Ru NNP-pincer complex 42. In addition, an analog Fe-pincer
complex was proposed [101]. The suggested mechanism runs via three catalytic
cycles, each with a direct hydride and ligand proton transfers. The cascade inter-
mediates are methyl formate, followed by formaldehyde and subsequently MeOH
as the product. The overall rate-determining step was found to be the formation of
the second MeOH molecule from methyl formate in the second cycle.
However, it should be kept in mind that the reduction of dimethyl carbonate to
methanol is uneconomic as the starting material with 1,000 US/t is much more
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 91

H
Pt Bu2
CO2
MeOH 2 MeOH MeOH 2 NHR2
RCH2NH2 N Ru CO
H2
NEt2 42
O O O O

H
H OMe MeO OMe RH2CHN OMe R2N NR2
Pt Bu2
2 H2 3 H2 3 H2 3 H2
N Ru CO
MeOH N
46 or 42 46 or 42 46

Scheme 4.12 Various ways for the indirect conversion of CO2 to MeOH published by Milstein

O
O cat. H2O omega
+ CO2 O O + CO2
HO OH process
H
H
N PPh2 47,H2, KOt Bu
Ru + MeOH Ref. [103]
HO OH
P CO
Ph2 Cl
47

Scheme 4.13 Another concept for the indirect production of MeOH from CO2 demonstrated
by Ding

expensive than MeOH (400 US/t) [102]. Therefore, Ding et al. developed another
indirect route for the MeOH production [102]. The idea is based on the omega
process where ethylene carbonate, synthesized from CO2 and ethylene oxide, is
hydrolyzed to ethylene glycol and CO2 (Scheme 4.13).
Instead of hydrolyzing the organic carbonate, this intermediate can be hydroge-
nated with the Ru PNP-pincer complex 47 in the presence of KOtBu to give
ethylene glycol and MeOH. This opens a nice way to get two bulk chemicals at
the same time from reacting CO2 with ethylene oxide. With TON up to 87,000 and
TOF of 1,200 h1, high activity could be observed at almost quantitative yields. In
addition, deuterated methanol could be obtained in 99 % yield by deuteration of the
sterically hindered tetramethyl ethylene carbonate with D2. Furthermore, the cata-
lyst is able to depolymerize the widely used material poly(propylene carbonate) by
hydrogenating it to MeOH and the corresponding diol. Therefore, a recycling of this
material to valuable products is possible [103].
An unusual way to gain methanol is the formic acid disproportionation (Eq. 4.3).
Here, formic acid is both substrate and hydrogen source. In a HCOOH/water
mixture at 60  C, the molecular iridium catalyst [Cp*Ir(bpy)(H2O)](OTf)2 converts
formic acid to methanol with a selectivity up to 12 % [104]. The selectivity
increases with higher HCOOH concentration and lower temperature. Additional
92 C. Ziebart and M. Beller

H2 suppresses the competing formate decomposition to CO2 and H2. As a side


product methyl formate was observed through HCOOH esterification. Deuteration
experiments indicate that formaldehyde is an intermediate of the catalytic cycle
which is then hydrogenated to MeOH by HCOOH releasing CO2.

3HCOOHðaqÞ ! CH3 OHðaqÞ þ H2 OðlÞ þ 2CO2ðgÞ ΔG 298 ¼ 11:9 kcal mol1 ð4:3Þ

An interesting approach is the use of frustrated Lewis pairs which are capable of
heterolytic H2 activation [105]. Ashley and O’Hare applied TMP + B(C6F5)3,
forming a unique formato borate complex [TMPH]-[HCO2B(C6F5)3] with CO2
under H2 atmosphere (1–2 atm). After 6 days at 160  C, H3COB(C6F5)2 was formed
and 24 % yield of methanol was observed after distillation. However, the rendering
of the catalytic cycle is still unachieved [106].

4.3.2 Catalytic Reduction with Other Hydride Reagents

There is a bunch of catalysts capable of reducing CO2 using other hydride reagents
such as aminoboranes, Hboranes and Hsilanes. For example, the group of Stephan
presented a 1:2 mixture of PMes3 and AlX3 reacting with CO2 to the corresponding
phosphonium-carboxylate adduct. This intermediate then dehydrogenates amino-
boranes at RT to give phosphonium methoxy aluminate in less than 15 min.
Quenching with H2O leads to 37–51 % yield of methanol [107].
The first catalytic metal-free hydrosilylation of CO2 to MeOH was reported in
2009 by Zhang and Ying [108]. Using the NHC catalyst 49 under CO2 atmosphere,
first imidazolium carboxylate is formed which reacted then with diphenylsilane at
RT to give CH3OSiMe3. Final quenching with water gives MeOH in 90 % yield
based on the used Hsilane amount with TON of 1,840 and TOF of 25 h1.
Therefore, it is competitive with known Ru catalysts lacking their typical draw-
backs of moisture and air sensitivity. This system is even active under dried air. The
first catalytic hydroboration of CO2 to the methoxide level was reported by Guan,
applying the nickel catalyst 48 [109] in the presence of HBcat. Hydrolysis gives
MeOH with TON and TOF of 495 h1 (based on B-H).
Comparing both systems by studying the mechanism quantum mechanically,
Wang proposed that the most favorable pathway for both contains analog, exper-
imentally detected intermediates like formoxy, acetal, and methoxide species [110,
111]. In addition, he predicted formaldehyde to be another intermediate in the
catalytic cycle (Scheme 4.14). Both catalysts accelerate the hydride transfer in
comparison to uncatalyzed reactions, but the mechanism of the hydride transfer is
different. In the Ni/borane system, the Ni catalyst 48 acts as a shuttle, transferring
the hydride from the HBcat to the C-O bonds. In the first step, CO2 inserts into the
[Ni]-H bond (50) forming Ni-formate. This complex reacts with catecholborane
HBcat (51) to give formyl borate and [Ni]-H catalyst 48. The catalyst then activates
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 93

H
t Bu2P Ni Pt Bu2
O O N N
Mes Mes

NHC
CO2 + 3 HBcat CH3OBcat + catBOBcat CO2 + 3 R3SiH CH3OSiR3 + R3SiOSiR3
48 49
H2O H2O
CH3OH CH3OH

+ HBcat + [Ni]H + NHC - NHC


HCOOBcat [Ni]OOCH CO2 52 HCOO[Si]
- [Ni]H + R3SiH
+ NHC*[Si]H
+ [Ni]H
50 - NHC
O
[Ni] N N
C O Mes Mes
[Ni]OCH2OBcat [Si]OCH2O[Si]
H
R Si H O
+ HBcat Bcat 52 C
O R R
- [Ni]H Bcat
H or O
O H H or
O H
- catBOBcat [Ni] - [Si]O[Si]
51 O [Ni] N N
Mes Mes

HCHO O O HCHO

R Si H + NHC*[Si]H
+ [Ni]H
R R - NHC
+ HBcat
- [Ni]H
CH3O[Ni] CH3OBcat MeOH H3CO[Si]
in H2O in H2O

Scheme 4.14 Comparing two catalytic systems based on the DFT studies by Wang for CO2 to
MeOH using hydroboration and hydrosilylation reagents

this product by forming the acetal species. The hydride transfer to [Ni] by a second
HBcat results in formaldehyde and di(boryl)ether production. Formaldehyde then
inserts into the [Ni]-H, yielding the corresponding methoxide complex. A third
HBcat closes the catalytic cycle by reforming the [Ni]-H catalyst and gaining the
final product CH3OBcat.
In contrast, the NHC catalyst 49 “only” activates the Si-H bond by pushing
electron density to the H atoms and simplifies this way the direct hydride transfer
from Hsilane to the C-O bond (52 top). However, the NHC catalyst 49 can react
with CO2 to the corresponding imidazolium carboxylate, too. This activated CO2
adduct can be attacked more easily by Hsilanes than unbounded CO2 (52 bottom).
It is noteworthy that the CO2 reduction catalyst and the hydride source must
match for achieving high activity. Exchanging the hydride source for both systems,
the reduction of CO2 to MeOH failed. The reasons are a high kinetic barrier for the
reaction of the Ni-formate with silanes and the formation of a thermodynamic
stable intermediate between NHC and Hboranes.
Further examples of a catalytic hydroboration with Ru catalysts were presented
by Sabo-Etienne using [RuH2(H2)2(PCy3)2] in combination with pinacolborane
[112, 113] and by the group of Stephan applying the precursor [Ru(PPh3)Cl2] and
the ligand N((CH2)2NHPiPr2)3 in the presence of HBpin [114].
94 C. Ziebart and M. Beller

The first metal-free organocatalyst for hydroboration of CO2 to MeOH was


published by Maron and Fontaine [115]. The organocatalyst 1-Bcat-2-PPh2-C6H4
reduces CO2 with 3 eq of BH3∙SMe2 to give subsequent MeOH with TON >3,000
and TOF >970 h1.

4.4 Other Homogeneous Catalytic Reductions of CO2


Using Hydride Reagents

Beside methanol, the full reduction of CO2 to methane is an attractive and actual
goal for homogeneous catalysis. However, so far homogeneous catalysts capable of
hydrogenating CO2 to methane with molecular H2 are still unknown (Eq. 4.4). In
the meantime, some efforts have been undertaken using less benign reductants such
as silanes.

CO2 þ 4 H2 ! CH4 þ 2 H2 O ð4:4Þ

Interestingly, Musgrave’s ab initio studies claim that even in the absence of a


metal, the reduction of CO2 with ammonia boranes via simultaneous hydride and
proton transfer should be possible involving the intermediates HCOOH and
hydrated formaldehyde to give finally methanol [116]. Unfortunately, ammonia
boranes decompose easily in the presence of acid to H2 and NH2BH2. Unless this
problem will be solved, the reduction will stop at the HCOOH step. The group of
Wehmschulte showed that strong Lewis acids such as [Et2Al]+ can catalyze the
hydrosilylation of CO2 to methane. However, after three cycles the activity
decreased, likely caused by side reactions with the solvent C6D6 [117].
Combining borane complexes or additional borane with hydrosilylation reagents
was the key to success for improved catalytic turnover numbers. All active systems
have in common that different kind of metal complexes or frustrated Lewis pairs
(FLP) are needed to activate the CO2 to form the corresponding formate complexes.
The following reduction steps are believed to be catalyzed by species [R3Si-H∙∙∙BR3]
involving the intermediates bis(silyl)acetal, methyl silyl ether, and finally methane.
Active co-catalysts are presented in Scheme 4.15.
Activities and selectivities were low for the systems using 53 [118], 54 [119],
and 55 [120], whereas a good TON of 2,156 was achieved in16 h by Turculet using
the Pt-pincer borane complex 56 with Et3SiH at 65  C and 1 atm of CO2 [121].
A high TON of 8,300 after 72 h at RT was published by Brookhart applying the
Ir+ PCP-pincer complex 57 [122] in the presence of Me2PhSiH.
Finally, it is worth mentioning that carbon dioxide can be used as methylation
reagent for amines in the presence of silanes. Our group [123] as well as the group
of Cantat [124] realized the reduction of CO2 with a simultaneous C-N bond
formation. With this method it is possible to methylate primary amines with CO2
as C1 feedstock.
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 95

PhH2C CH2Ph
t Bu O t Bu
Zr O
B(C6F5)3
O Me NH
[Cp*2Sc][HB(C6F5)3]
TMP
Me Me
53 54 55
B(C6F5)3

PCy2
O
t Bu2P Ir Pt Bu2
MeSi Pt HB(C6F5)3
H
O O
PCy2
B(C6F5)4
56 57

Scheme 4.15 Various co-catalysts for the [R3Si-H∙∙∙BR3] catalyzed reduction of CO2 to CH4

4.5 Concluding Remarks

In the last decade, several improved organometallic catalysts for the hydrogenation
of CO2 to formic acid derivatives have been reported. Many of them show good
activity even under ambient conditions. In addition, interesting concepts have been
developed, combining both formic acid product separation and catalyst recycling.
However, for industrial applications the activities and long-term stabilities of the
catalysts still have to be improved. Meanwhile, also catalyst systems have been
presented for hydrogen storage using carbon dioxide.
While non-precious metals such as Fe and Co were rarely known for the
homogeneous CO2 hydrogenation before 2009, much progress was achieved in
this field in the last 3 years.
Very recently, also first promising systems capable of converting CO2 to methanol
became available opening new ways to a future methanol economy. On the other
hand, the homogeneously catalyzed reduction of carbon dioxide to CH4 using
molecular hydrogen is still not known and more efforts are needed in this area. At
this point it should be mentioned that related reductions of carbon dioxide using
boranes or hydrosilylation reagents are interesting in order to understand the mech-
anism of such processes. Nevertheless, it should be clear that they will not be of
preparative use due to price, sensitivity of the reagents, and waste generated.
What will be the future challenges in homogeneous CO2 reductions? Obviously,
for industrial implementations even better catalysts are needed. With regard to basic
science, the hydrogenation of carbon dioxide with simultaneous C-C bond formation
represents a grand challenge. As an example, the low temperature and selective
Fischer-Tropsch reaction can be considered. Furthermore, reducing carbon dioxide
directly to Co is of both industrial and academic interest. Such transformations would
allow using carbon dioxide instead of toxic carbon monoxide. The key tools to achieve
these goals will be catalysis and organometallic chemistry.
96 C. Ziebart and M. Beller

References

1. Cokoja M, Bruckmeier C, Rieger B, Herrmann WA, Kühn FE (2011) Umwandlung von


Kohlendioxid mit Übergangsmetall-Homogenkatalysatoren: eine molekulare Lösung für ein
globales Problem? Angew Chem 123(37):8662–8690. doi:10.1002/ange.201102010
2. Peters M, Köhler B, Kuckshinrichs W, Leitner W, Markewitz P, Müller TE (2011) Chemical
technologies for exploiting and recycling carbon dioxide into the value chain. ChemSusChem
4(9):1216–1240. doi:10.1002/cssc.201000447
3. Geider RJ, Delucia EH, Falkowski PG, Finzi AC, Grime JP, Grace J, Kana TM, La Roche J,
Long SP, Osborne BA, Platt T, Prentice IC, Raven JA, Schlesinger WH, Smetacek V,
Stuart V, Sathyendranath S, Thomas RB, Vogelmann TC, Williams P, Woodward FI
(2001) Primary productivity of planet earth: biological determinants and physical constraints
in terrestrial and aquatic habitats. Glob Chang Biol 7(8):849–882. doi:10.1046/j.1365-2486.
2001.00448.x
4. Appel AM, Bercaw JE, Bocarsly AB, Dobbek H, DuBois DL, Dupuis M, Ferry JG, Fujita E,
Hille R, Kenis PJA, Kerfeld CA, Morris RH, Peden CHF, Portis AR, Ragsdale SW,
Rauchfuss TB, Reek JNH, Seefeldt LC, Thauer RK, Waldrop GL (2013) Frontiers, opportu-
nities, and challenges in biochemical and chemical catalysis of CO2 fixation. Chem Rev.
doi:10.1021/cr300463y
5. Mikkelsen M, Jorgensen M, Krebs FC (2010) The teraton challenge. A review of fixation and
transformation of carbon dioxide. Energy Environ Sci 3(1):43–81
6. Bertau M, Offermanns H, Menges G, Keim W, Effenberger FX (2010) Methanol findet zu
wenig Beachtung als Kraftstoff und Chemierohstoff der Zukunft [Methanol needs more
attention as a fuel and raw material for the future]. Chem Ing Tech 82(12):2055–2058.
doi:10.1002/cite.201000159
7. Sakakura T, Choi J-C, Yasuda H (2007) Transformation of carbon dioxide. Chem Rev 107
(6):2365–2387. doi:10.1021/cr068357u
8. Olah GA, Prakash GKS, Goeppert A (2011) Anthropogenic chemical carbon cycle for a
sustainable future. J Am Chem Soc 133(33):12881–12898. doi:10.1021/ja202642y
9. Olah GA (2013) Towards oil independence through renewable methanol chemistry. Angew
Chem Int Ed 52(1):104–107. doi:10.1002/anie.201204995
10. Haggin J (1994) New processes target methanol production, off-gas cleaning. Chem Eng
News 72:29–31
11. Goehna H, Koenig P (1994) Producing methanol from CO2. ChemTech 24:36–39
12. Jessop PG, Ikariya T, Noyori R (1995) Homogeneous hydrogenation of carbon dioxide.
Chem Rev 95(2):259–272. doi:10.1021/cr00034a001
13. Leitner W (1995) Carbon dioxide as a raw material: the synthesis of formic acid and its
derivatives from CO2. Angew Chem Int Ed 34(20):2207–2221. doi:10.1002/anie.199522071
14. Jessop PG, Joó F, Tai C-C (2004) Recent advances in the homogeneous hydrogenation of
carbon dioxide. Coord Chem Rev 248(21–24):2425–2442. doi:10.1016/j.ccr.2004.05.019
15. de Vries JG, Elsevier CJ, Jessop PG (2007) 17 homogeneous hydrogenation of carbon
dioxide. Handb Homogen Hydrogen 1:489–511
16. Rothenberg G (2008) Introduction. In: Catalysis. Wiley-VCH Verlag GmbH & Co. KGaA,
Weinheim, pp 1–38. doi:10.1002/9783527621866.ch1
17. Olah GA (2005) Beyond oil and gas: the methanol economy. Angew Chem Int Ed
44(18):2636–2639. doi:10.1002/anie.200462121
18. Munshi P, Main AD, Linehan JC, Tai C-C, Jessop PG (2002) Hydrogenation of carbon
dioxide catalyzed by ruthenium trimethylphosphine complexes: the accelerating effect of
certain alcohols and amines. J Am Chem Soc 124(27):7963–7971. doi:10.1021/ja0167856
19. Y-y O, Matsunaga T, Nakao Y, Sato H, Sakaki S (2005) Ruthenium(II)-catalyzed hydroge-
nation of carbon dioxide to formic acid. Theoretical study of real catalyst, ligand effects, and
solvation effects. J Am Chem Soc 127(11):4021–4032. doi:10.1021/ja043697n
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 97

20. Y-y O, Nakao Y, Sato H, Sakaki S (2006) Ruthenium(II)-catalyzed hydrogenation of carbon


dioxide to formic acid. Theoretical study of significant acceleration by water molecules.
Organometallics 25(14):3352–3363. doi:10.1021/om060307s
21. Getty AD, Tai C-C, Linehan JC, Jessop PG, Olmstead MM, Rheingold AL (2009) Hydroge-
nation of carbon dioxide catalyzed by ruthenium trimethylphosphine complexes: a mechanistic
investigation using high-pressure NMR spectroscopy. Organometallics 28(18):5466–5477.
doi:10.1021/om900128s
22. Gowrisankar S, Federsel C, Neumann H, Ziebart C, Jackstell R, Spannenberg A, Beller M
(2013) Synthesis of stable phosphomide ligands and their use in Ru-catalyzed hydrogenations
of bicarbonate and related substrates. ChemSusChem 6(1):85–91. doi:10.1002/cssc.201200732
23. Muller K, Sun Y, Thiel WR (2013) Ruthenium(II)–phosphite complexes as catalysts for the
hydrogenation of carbon dioxide. ChemCatChem 5(6):1340–1343. doi:10.1002/cctc.
201200818
24. Zhao G, Joó F (2011) Free formic acid by hydrogenation of carbon dioxide in sodium formate
solutions. Catal Commun 14(1):74–76. doi:10.1016/j.catcom.2011.07.017
25. Johnson TC, Morris DJ, Wills M (2010) Hydrogen generation from formic acid and alcohols
using homogeneous catalysts. Chem Soc Rev 39(1):81–88
26. Grasemann M, Laurenczy G (2012) Formic acid as a hydrogen source – recent developments
and future trends. Energy Environ Sci 5(8):8171–8181. doi:10.1039/C2EE21928J
27. Loges B, Boddien A, Gärtner F, Junge H, Beller M (2010) Catalytic generation of hydrogen
from formic acid and its derivatives: useful hydrogen storage materials. Top Catal 53:902–914
28. Laurenczy G, Dalebrook AF, Gan W, Grasemann M, Moret S (2013) Hydrogen storage –
beyond conventional methods. Chem Commun. doi:10.1039/C3CC43836H
29. Graetz J (2009) New approaches to hydrogen storage. Chem Soc Rev 38(1):73–82. doi:10.
1039/B718842K
30. Jena P (2011) Materials for hydrogen storage: past, present, and future. J Phys Chem Lett 2
(3):206–211. doi:10.1021/jz1015372
31. Eberle U, Felderhoff M, Schüth F (2009) Chemical and physical solutions for hydrogen
storage. Angew Chem Int Ed 48(36):6608–6630. doi:10.1002/anie.200806293
32. Papp G, Csorba J, Laurenczy G, Joó F (2011) A charge/discharge device for chemical
hydrogen storage and generation. Angew Chem 123(44):10617–10619. doi:10.1002/ange.
201104951
33. Boddien A, Gärtner F, Federsel C, Sponholz P, Mellmann D, Jackstell R, Junge H, Beller M
(2011) Kohlenstoffdioxid-neutrale Wasserstoffspeicherung basierend auf Bicarbonaten und
Formiaten. Angew Chem 123:6535–6538. doi:10.1002/ange.201101995
34. Federsel C, Jackstell R, Boddien A, Laurenczy G, Beller M (2010) Ruthenium-catalyzed
hydrogenation of bicarbonate in water. ChemSusChem 3(9):1048–1050. doi:10.1002/cssc.
201000151
35. Boddien A, Federsel C, Sponholz P, Mellmann D, Jackstell R, Junge H, Laurenczy G, Beller M
(2012) Towards the development of a hydrogen battery. Energy Environ Sci 5(10):8907–8911.
doi:10.1039/c2ee22043a
36. Urakawa A, Iannuzzi M, Hutter J, Baiker A (2007) Towards a rational design of ruthenium
CO2 hydrogenation catalysts by ab initio metadynamics. Chem Eur J 13(24):6828–6840.
doi:10.1002/chem.200700254
37. Urakawa A, Jutz F, Laurenczy G, Baiker A (2007) Carbon dioxide hydrogenation catalyzed
by a ruthenium dihydride: a DFT and high-pressure spectroscopic investigation. Chem Eur J
13(14):3886–3899. doi:10.1002/chem.200601339
38. Dietrich J, Schindler S (2008) Kinetic studies on the hydrogenation of carbon dioxide to
formic acid using a rhodium complex as catalyst. Z Anorg Allg Chem 634(14):2487–2494.
doi:10.1002/zaac.200800352
39. Yasaka Y, Wakai C, Matubayasi N, Nakahara M (2010) Controlling the equilibrium of
formic acid with hydrogen and carbon dioxide using ionic liquid. J Phys Chem A 114
(10):3510–3515. doi:10.1021/jp908174s
98 C. Ziebart and M. Beller

40. Zhang Z, Xie Y, Li W, Hu S, Song J, Jiang T, Han B (2008) Hydrogenation of carbon dioxide
is promoted by a task-specific ionic liquid. Angew Chem Int Ed 47(6):1127–1129. doi:10.
1002/anie.200704487
41. Zhang Z, Hu S, Song J, Li W, Yang G, Han B (2009) Hydrogenation of CO2 to formic acid
promoted by a diamine-functionalized ionic liquid. ChemSusChem 2(3):234–238. doi:10.
1002/cssc.200800252
42. Wesselbaum S, Hintermair U, Leitner W (2012) Continuous-flow hydrogenation of carbon
dioxide to pure formic acid using an integrated scCO2 process with immobilized catalyst and
base. Angew Chem 124:8713–8716. doi:10.1002/ange.201203185
43. Schaub T, Paciello RA (2011) A process for the synthesis of formic acid by CO2 hydroge-
nation: thermodynamic aspects and the role of CO. Angew Chem Int Ed 50(32):7278–7282.
doi:10.1002/anie.201101292
44. Ogo S, Kabe R, Hayashi H, Harada R, Fukuzumi S (2006) Mechanistic investigation of CO2
hydrogenation by Ru(II) and Ir(III) aqua complexes under acidic conditions: two catalytic
systems differing in the nature of the rate determining step. Dalton Trans 39:4657–4663.
doi:10.1039/b607993h
45. Himeda Y, Onozawa-Komatsuzaki N, Sugihara H, Kasuga K (2007) Simultaneous tuning of
activity and water solubility of complex catalysts by acid–base equilibrium of ligands for
conversion of carbon dioxide. Organometallics 26(3):702–712. doi:10.1021/om060899e712
46. Himeda Y, Miyazawa S, Hirose T (2011) Interconversion between formic acid and H2/CO2
using rhodium and ruthenium catalysts for CO2 fixation and H2 storage. ChemSusChem 4
(4):487–493. doi:10.1002/cssc.201000327
47. Himeda Y (2007) Conversion of CO2 into formate by homogeneously catalyzed hydrogena-
tion in water: tuning catalytic activity and water solubility through the acid–base equilibrium
of the ligand. Eur J Inorg Chem 25:3927–3941. doi:10.1002/ejic.200700494
48. Wang W-H, Hull JF, Muckerman JT, Fujita E, Himeda Y (2012) Second-coordination-sphere
and electronic effects enhance iridium(III)-catalyzed homogeneous hydrogenation of carbon
dioxide in water near ambient temperature and pressure. Energy Environ Sci 5:7923–7926
49. Wang W-H, Muckerman JT, Fujita E, Himeda Y (2013) Mechanistic insight through factors
controlling effective hydrogenation of CO2 catalyzed by bioinspired proton-responsive
iridium(III) complexes. ACS Catal 3(5):856–860. doi:10.1021/cs400172j
50. Hull JF, Himeda Y, Wang W-H, Hashiguchi B, Periana R, Szalda DJ, Muckerman JT, Fujita
E (2012) Reversible hydrogen storage using CO2 and a proton-switchable iridium catalyst in
aqueous media under mild temperatures and pressures. Nat Chem 4(5):383–388
51. Maenaka Y, Suenobu T, Fukuzumi S (2012) Catalytic interconversion between hydrogen and
formic acid at ambient temperature and pressure. Energy Environ Sci 5(6):7360–7367.
doi:10.1039/c2ee03315a
52. Fukuzumi S, Suenobu T (2013) Hydrogen storage and evolution catalysed by metal hydride
complexes. Dalton Trans 42(1):18–28. doi:10.1039/c2dt31823g
53. Muller K, Sun Y, Heimermann A, Menges F, Niedner-Schatteburg G, van Wüllen C, Thiel
WR (2013) Structure–reactivity relationships in the hydrogenation of carbon dioxide with
ruthenium complexes bearing pyridinylazolato ligands. Chem Eur J 19:7825–7834. doi:10.
1002/chem.201204199
54. Thai T-T, Therrien B, Suss-Fink G (2009) Arene ruthenium oxinato complexes: synthesis,
molecular structure and catalytic activity for the hydrogenation of carbon dioxide in aqueous
solution. J Organomet Chem 694(25):3973–3981. doi:10.1016/j.jorganchem.2009.09.008
55. Sanz S, Azua A, Peris E (2010) ‘([small eta]6-arene)Ru(bis-NHC)’ complexes for the
reduction of CO2 to formate with hydrogen and by transfer hydrogenation with iPrOH.
Dalton Trans 39(27):6339–6343. doi:10.1039/C003220D
56. Azua A, Sanz S, Peris E (2011) Water-soluble IrIII N-heterocyclic carbene based catalysts for
the reduction of CO2 to formate by transfer hydrogenation and the deuteration of aryl amines
in water. Chem Eur J 17(14):3963–3967. doi:10.1002/chem.201002907
57. Albrecht M, van Koten G (2001) Platinum group organometallics based on “pincer” complexes:
sensors, switches, and catalysts. Angew Chem Int Ed 40(20):3750–3781. doi:10.1002/1521-
3773(20011015)40:20<3750::aid-anie3750>3.0.co;2-6
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 99

58. van der Boom ME, Milstein D (2003) Cyclometalated phosphine-based pincer complexes:
mechanistic insight in catalysis, coordination, and bond activation. Chem Rev
103(5):1759–1792. doi:10.1021/cr960118r
59. Grützmacher H (2008) Cooperating ligands in catalysis. Angew Chem Int Ed
47(10):1814–1818. doi:10.1002/anie.200704654
60. van der Vlugt JI, Reek JNH (2009) Neutral tridentate PNP ligands and their hybrid analogues:
versatile non-innocent scaffolds for homogeneous catalysis. Angew Chem Int Ed
48(47):8832–8846. doi:10.1002/anie.200903193
61. Musa S, Shaposhnikov I, Cohen S, Gelman D (2011) Ligand–metal cooperation in PCP
pincer complexes: rational design and catalytic activity in acceptorless dehydrogenation of
alcohols. Angew Chem 123(15):3595–3599. doi:10.1002/ange.201007367
62. Federsel C, Jackstell R, Beller M (2010) State-of-the-art catalysts for hydrogenation of
carbon dioxide. Angew Chem Int Ed 49(36):6254–6257. doi:10.1002/anie.201000533
63. Nozaki K, Tanaka R, Yamashita M (2009) Catalytic hydrogenation of carbon dioxide using Ir
(III) – pincer complexes. J Am Chem Soc 131(40):14168–14169. doi:10.1021/ja903574e
64. Yang X (2011) Hydrogenation of carbon dioxide catalyzed by PNP pincer iridium, iron, and
cobalt complexes: a computational design of base metal catalysts. ACS Catal 1:849–854.
doi:10.1021/cs2000329
65. Ahlquist MSG (2010) Iridium catalyzed hydrogenation of CO2 under basic conditions-
mechanistic insight from theory. J Mol Catal A Chem 324(1–2):3–8. doi:10.1016/j.
molcata.2010.02.018
66. Tanaka R, Yamashita M, Chung LW, Morokuma K, Nozaki K (2011) Mechanistic studies on
the reversible hydrogenation of carbon dioxide catalyzed by an Ir-PNP complex. Organome-
tallics 30(24):6742–6750. doi:10.1021/om2010172
67. Schmeier TJ, Dobereiner GE, Crabtree RH, Hazari N (2011) Secondary coordination sphere
interactions facilitate the insertion step in an iridium(III) CO2 reduction catalyst. J Am Chem
Soc 133(24):9274–9277. doi:10.1021/ja2035514
68. Vogt M, Rivada-Wheelaghan O, Iron MA, Leitus G, Diskin-Posner Y, Shimon LJW,
Ben-David Y, Milstein D (2013) Anionic Nickel(II) complexes with doubly deprotonated
PNP pincer-type ligands and their reactivity toward CO2. Organometallics 32(1):300–308.
doi:10.1021/om3010838
69. Khaskin E, Diskin-Posner Y, Weiner L, Leitus G, Milstein D (2013) Formal loss of an H radical
by a cobalt complex via metal-ligand cooperation. Chem Commun 49(27):2771–2773.
doi:10.1039/C3CC39049G
70. Tai C-C, Chang T, Roller B, Jessop PG (2003) High-pressure combinatorial screening of
homogeneous catalysts: hydrogenation of carbon dioxide. Inorg Chem 42(23):7340–7341.
doi:10.1021/ic034881x
71. Langer R, Diskin-Posner Y, Leitus G, Shimon LJW, Ben-David Y, Milstein D (2011)
Low-pressure hydrogenation of carbon dioxide catalyzed by an iron pincer complex
exhibiting noble metal activity. Angew Chem Int Ed 50(42):9948–9952. doi:10.1002/anie.
201104542
72. Federsel C, Boddien A, Jackstell R, Jennerjahn R, Dyson PJ, Scopelliti R, Laurenczy G, Beller
M (2010) A well-defined iron catalyst for the reduction of bicarbonates and carbon dioxide
to formates, alkyl formates, and formamides. Angew Chem Int Ed 49(50):9777–9780.
doi:10.1002/anie.201004263
73. Federsel C, Ziebart C, Jackstell R, Baumann W, Beller M (2012) Catalytic hydrogenation of
carbon dioxide and bicarbonates with a well-defined cobalt dihydrogen complex. Chem Eur J
18(1):72–75. doi:10.1002/chem.201101343
74. Ziebart C, Federsel C, Anbarasan P, Jackstell R, Baumann W, Spannenberg A, Beller M
(2012) Well-defined iron catalyst for improved hydrogenation of carbon dioxide and bicar-
bonate. J Am Chem Soc 134(51):20701–20704. doi:10.1021/ja307924a
75. Bianchini C, Peruzzini M, Zanobini F (1988) An exceptionally stable cis-(hydride)([eta]2-
dihydrogen) complex of iron. J Organomet Chem 354(2):C19–C22. doi:10.1016/0022-328x
(88)87057-8
100 C. Ziebart and M. Beller

76. Jeletic MS, Mock MT, Appel AM, Linehan JC (2013) A cobalt-based catalyst for the
hydrogenation of CO2 under ambient conditions. J Am Chem Soc. doi:10.1021/ja406601v
77. Jansen A, Pitter S (2004) Homogeneously catalysed reduction of carbon dioxide with silanes:
a study on solvent and ligand effects and catalyst recycling. J Mol Catal A Chem 217
(1–2):41–45, http://dx.doi.org/10.1016/j.molcata.2004.03.041
78. Deglmann P, Ember E, Hofmann P, Pitter S, Walter O (2007) Experimental and theoretical
investigations on the catalytic hydrosilylation of carbon dioxide with ruthenium nitrile
complexes. Chem Eur J 13(10):2864–2879. doi:10.1002/chem.200600396
79. Motokura K, Kashiwame D, Takahashi N, Miyaji A, Baba T (2013) Highly active and
selective catalysis of copper diphosphine complexes for the transformation of carbon dioxide
into silyl formate. Chem Eur J 19(30):10030–10037. doi:10.1002/chem.201300935
80. Lalrempuia R, Iglesias M, Polo V, Sanz Miguel PJ, Fernández-Alvarez FJ, Pérez-Torrente JJ,
Oro LA (2012) Effective fixation of CO2 by iridium-catalyzed hydrosilylation. Angew Chem
Int Ed 51(51):12824–12827. doi:10.1002/anie.201206165
81. Sattler W, Parkin G (2012) Zinc catalysts for on-demand hydrogen generation and carbon
dioxide functionalization. J Am Chem Soc 134(42):17462–17465. doi:10.1021/ja308500s
82. Motokura K, Kashiwame D, Miyaji A, Baba T (2012) Copper-catalyzed formic acid synthesis
from CO2 with hydrosilanes and H2O. Org Lett 14(10):2642–2645. doi:10.1021/ol301034j
83. Zhang L, Cheng J, Hou Z (2013) Highly efficient catalytic hydrosilylation of carbon dioxide
by an N-heterocyclic carbene copper catalyst. Chem Commun 49(42):4782–4784. doi:10.
1039/C3CC41838C
84. Itagaki S, Yamaguchi K, Mizuno N (2013) Catalytic synthesis of silyl formates with 1 atm of
CO2 and their utilization for synthesis of formyl compounds and formic acid. J Mol Catal A
Chem 366:347–352, http://dx.doi.org/10.1016/j.molcata.2012.10.014
85. Schäfer A, Saak W, Haase D, Müller T (2012) Silyl cation mediated conversion of CO2 into
benzoic acid, formic acid, and methanol. Angew Chem Int Ed 51(12):2981–2984. doi:10.
1002/anie.201107958
86. Miller AJM, Labinger JA, Bercaw JE (2011) Trialkylborane-assisted CO2 reduction by late
transition metal hydrides. Organometallics 30(16):4308–4314. doi:10.1021/om200364w
87. Shintani R, Nozaki K (2013) Copper-catalyzed hydroboration of carbon dioxide. Organome-
tallics 32(8):2459–2462. doi:10.1021/om400175h
88. Stephan DW (2013) Catalysis: a step closer to a methanol economy. Nature 495(7439):54–55
89. Nielsen M, Alberico E, Baumann W, Drexler H-J, Junge H, Gladiali S, Beller M (2013)
Low-temperature aqueous-phase methanol dehydrogenation to hydrogen and carbon dioxide.
Nature 495(7439):85–89
90. (a) Surumpudi S, Narayanan SR, Vamos E, Frank HA, Halpert G, Prakash GKS, Olah GA,
US Patent 6,248,460, 2001; (b) Surumpudi S, Narayanan SR, Vamos E, Frank HA, Halpert G,
Olah GA, Prakash GKS, US Patent 6,444,343, 2002
91. Chang CD (1997) In: Ertl G, Knözinger H, Weitkamp J (eds) Handbook of heterogeneous
catalysis. VCH, Weinheim, p 1894 and references therein
92. Huff CA, Sanford MS (2011) Cascade catalysis for the homogeneous hydrogenation of CO2
to methanol. J Am Chem Soc 133(45):18122–18125. doi:10.1021/ja208760j
93. Huff CA, Kampf JW, Sanford MS (2012) Role of a noninnocent pincer ligand in
the activation of CO2 at (PNN)Ru(H)(CO). Organometallics 31(13):4643–4645.
doi:10.1021/om300403b
94. Vogt M, Gargir M, Iron MA, Diskin-Posner Y, Ben-David Y, Milstein D (2012) A new mode of
activation of CO2 by metal–ligand cooperation with reversible C–C and M–O bond formation
at ambient temperature. Chem Eur J 18(30):9194–9197. doi:10.1002/chem.201201730
95. Langer J, Fabra MJ, Garcia-Orduna P, Lahoz FJ, Oro LA (2008) A C-H activation-CO2-
carboxylation reaction sequence mediated by an ‘Iridium(dppm)’ species. Formation of the
anionic ligand (Ph2P)2C-COOH. Chem Commun 39:4822–4824
96. Wesselbaum S, vom Stein T, Klankermayer J, Leitner W (2012) Hydrogenation of carbon
dioxide to methanol by using a homogeneous ruthenium–phosphine catalyst. Angew Chem
124:7617–7620. doi:10.1002/ange.201202320
4 Hydrogenation and Related Reductions of Carbon Dioxide with Molecular Catalysts 101

97. Balaraman E, Gunanathan C, Zhang J, Shimon LJW, Milstein D (2011) Efficient hydroge-
nation of organic carbonates, carbamates and formates indicates alternative routes to meth-
anol based on CO2 and CO. Nat Chem 3(8):609–614
98. Balaraman E, Ben-David Y, Milstein D (2011) Unprecedented catalytic hydrogenation of
urea derivatives to amines and methanol. Angew Chem Int Ed 50(49):11702–11705.
doi:10.1002/anie.201106612
99. Dixneuf PH (2011) Bifunctional catalysis: a bridge from CO2 to methanol. Nat Chem
3(8):578–579
100. Choudhury J (2012) New strategies for CO2-to-methanol conversion. ChemCatChem
4:609–611. doi:10.1002/cctc.201100495
101. Yang X (2012) Metal hydride and ligand proton transfer mechanism for the hydrogenation of
dimethyl carbonate to methanol catalyzed by a pincer ruthenium complex. ACS Catal
2(6):964–970. doi:10.1021/cs3000683
102. Han Z, Rong L, Wu J, Zhang L, Wang Z, Ding K (2012) Catalytic hydrogenation of cyclic
carbonates: a practical approach from CO2 and epoxides to methanol and diols. Angew Chem
Int Ed 51(52):13041–13045. doi:10.1002/anie.201207781
103. Li Y, Junge K, Beller M (2013) Improving the efficiency of the hydrogenation of carbonates
and carbon dioxide to methanol. ChemCatChem 5(5):1072–1074. doi:10.1002/cctc.
201300013
104. Miller AJM, Heinekey DM, Mayer JM, Goldberg KI (2013) Catalytic disproportionation of
formic acid to generate methanol. Angew Chem 125(14):4073–4076. doi:10.1002/ange.
201208470
105. Stephan DW, Erker G (2010) Frustrated Lewis pairs: metal-free hydrogen activation and
more. Angew Chem Int Ed 49(1):46–76. doi:10.1002/anie.200903708
106. Ashley AE, Thompson AL, O’Hare D (2009) Non-metal-mediated homogeneous hydroge-
nation of CO2 to CH3OH. Angew Chem Int Ed 48(52):9839–9843. doi:10.1002/anie.
200905466
107. Ménard G, Stephan DW (2010) Room temperature reduction of CO2 to methanol by Al-based
frustrated Lewis pairs and ammonia borane. J Am Chem Soc 132(6):1796–1797.
doi:10.1021/ja9104792
108. Riduan SN, Zhang Y, Ying JY (2009) Conversion of carbon dioxide into methanol with
silanes over N-heterocyclic carbene catalysts. Angew Chem Int Ed 48(18):3322–3325.
doi:10.1002/anie.200806058
109. Chakraborty S, Zhang J, Krause JA, Guan H (2010) An efficient nickel catalyst for the
reduction of carbon dioxide with a borane. J Am Chem Soc 132(26):8872–8873. doi:10.1021/
ja103982t
110. Huang F, Lu G, Zhao L, Li H, Wang Z-X (2010) The catalytic role of N-heterocyclic carbene
in a metal-free conversion of carbon dioxide into methanol: a computational mechanism
study. J Am Chem Soc 132(35):12388–12396. doi:10.1021/ja103531z
111. Huang F, Zhang C, Jiang J, Wang Z-X, Guan H (2011) How does the nickel pincer complex
catalyze the conversion of CO2 to a methanol derivative? A computational mechanistic study.
Inorg Chem 50(8):3816–3825. doi:10.1021/ic200221a
112. Bontemps S, Vendier L, Sabo-Etienne S (2012) Borane-mediated carbon dioxide reduction at
ruthenium: formation of C1 and C2 compounds. Angew Chem Int Ed 51(7):1671–1674.
doi:10.1002/anie.201107352
113. Bontemps S, Sabo-Etienne S (2013) Trapping formaldehyde in the homogeneous catalytic
reduction of carbon dioxide. Angew Chem Int Ed. doi:10.1002/anie.201304025
114. Sgro MJ, Stephan DW (2012) Frustrated Lewis pair inspired carbon dioxide reduction
by a ruthenium tris(aminophosphine) complex. Angew Chem 124(45):11505–11507.
doi:10.1002/ange.201205741
115. Courtemanche M-A, Légaré M-A, Maron L, Fontaine F-G (2013) A highly active
phosphine–borane organocatalyst for the reduction of CO2 to methanol using hydroboranes.
J Am Chem Soc 135(25):9326–9329. doi:10.1021/ja404585p
102 C. Ziebart and M. Beller

116. Zimmerman PM, Zhang Z, Musgrave CB (2010) Simultaneous two-hydrogen transfer as


a mechanism for efficient CO2 reduction. Inorg Chem 49(19):8724–8728. doi:10.1021/
ic100454z
117. Khandelwal M, Wehmschulte RJ (2012) Deoxygenative reduction of carbon dioxide to
methane, toluene, and diphenylmethane with [Et2Al]+ as catalyst. Angew Chem Int Ed
51(29):7323–7326. doi:10.1002/anie.201201282
118. Matsuo T, Kawaguchi H (2006) From carbon dioxide to methane: homogeneous reduction of
carbon dioxide with hydrosilanes catalyzed by zirconium – borane complexes. J Am Chem
Soc 128(38):12362–12363. doi:10.1021/ja0647250
119. Berkefeld A, Piers WE, Parvez M (2010) Tandem frustrated Lewis pair/Tris
(pentafluorophenyl)borane-catalyzed deoxygenative hydrosilylation of carbon dioxide.
J Am Chem Soc 132(31):10660–10661. doi:10.1021/ja105320c
120. Berkefeld A, Piers WE, Parvez M, Castro L, Maron L, Eisenstein O (2013) Decamethyls-
candocinium-hydrido-(perfluorophenyl)borate: fixation and tandem tris(perfluorophenyl)
borane catalysed deoxygenative hydrosilation of carbon dioxide. Chem Sci
4(5):2152–2162. doi:10.1039/C3SC50145K
121. Mitton SJ, Turculet L (2012) Mild reduction of carbon dioxide to methane with tertiary
silanes catalyzed by platinum and palladium silyl pincer complexes. Chem Eur J
18(48):15258–15262. doi:10.1002/chem.201203226
122. Park S, Bézier D, Brookhart M (2012) An efficient iridium catalyst for reduction of carbon
dioxide to methane with trialkylsilanes. J Am Chem Soc 134(28):11404–11407. doi:10.1021/
ja305318c
123. Li Y, Fang X, Junge K, Beller M (2013) A general catalytic methylation of amines using
carbon dioxide. Angew Chem Int Ed. doi:10.1002/anie.201301349
124. Jacquet O, Frogneux X, Das Neves Gomes C, Cantat T (2013) CO2 as a C1-building block for
the catalytic methylation of amines. Chem Sci 4(5):2127–2131. doi:10.1039/C3SC22240C
Chapter 5
Latest Advances in the Catalytic
Hydrogenation of Carbon Dioxide
to Methanol/Dimethylether

Francesco Arena, Giovanni Mezzatesta, Lorenzo Spadaro,


and Giuseppe Trunfio

5.1 Introduction

As world population growth places more demand on limited fossil fuels, the
discovery of renewable energy sources becomes crucial as part of the solution to
the impending energy and environmental dilemmas. Indeed, the continuously rising
levels of carbon dioxide in the atmosphere, expected to double current levels
(400 ppm) before 2050 [1], represent to date a major environmental threat,
owing to the unpredictable consequences of the greenhouse effect responsible for
the global warming and climate changes already in place [1–5]. This depends on the
fact that carbon dioxide is naturally recycled through photosynthesis processes,
using sun’s energy, in geological times.
Therefore, systematic exploitation of renewable energy sources and recycling
technologies are compulsory to attain a significant short-term cut of carbon dioxide
emissions. In this context a very viable solution for the abatement of the huge releases
from fossil fuel-burning power plants and other industrial flue gases, including cement
factories, consists of the chemical conversion of carbon dioxide to fuels alternative to
oil-derived ones, like methanol (MOH) and dimethylether (DME) [2–11].
Indeed, methanol and dimethylether are effective substitutes of LPG, gasoline and
gas oil, ensuring also considerably minor VOC and PM10 polluting impact than crude
oil-derived fuels [2, 5–7, 12]; with an average octane number of 100, the former has
excellent characteristics for spark-ignited engines, while DME (currently produced by
methanol dehydration) can well replace diesel in compression-ignited engines, because
of its high cetane number (55–60), and LPG for household purposes [2–5, 9, 10].

F. Arena (*) • L. Spadaro


DEEICE-UNIME + Instituto CNR-ITAE, Nicola Giordano,
Salita S. Lucia 31, 98125 Messina, Italy
e-mail: Francesco.Arena@unime.it
G. Mezzatesta • G. Trunfio
Department of Electronic Engineering, Industrial Chemistry and Engineering,
University of Messina, Messina, Italy

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 103
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_5,
© Springer-Verlag Berlin Heidelberg 2014
104 F. Arena et al.

In perspective, MOH and DME are advantageous to hydrogen as an energy carrier,


as the latter is explosive and volatile and has a very low energy density, also being a
prime material for synthetic hydrocarbon manufacture and chemical industry as well
[2, 4–9]. A remarkable advantage of MOH and DME also over biofuels is that the
former do not require shifting valuable agricultural and food resources to fuel
production. Therefore, an economy based on methanol as an energy carrier could
be renewable at the short timescales necessary for human needs.
Easily accessible from natural gas, MOH-based fuels would also meet the
strategic goal of reducing the oil dependence of Western countries, while the
exploitation of CO2-rich natural gas fields, coupled to lower costs of syngas
obtained by technologies alternative to steam reforming, like wet partial oxidation
(WPOX) or auto-thermal reforming (ATR), represents further remarkable eco-
nomic advantages of CO2-hydrogenation processes over current MOH–DME syn-
thesis technologies [5, 11–13].
Since many decades, methanol synthesis is industrially run on Cu–ZnO/Al2O3
catalysts, feeding syngas streams containing traces of CO2 [2, 4, 6, 14–16], though
methanol can be obtained via CO2 hydrogenation with rates and carbon utilisation
factors comparable to those of traditional route [17–19]. Despite mechanism and
active sites still being a matter of debate, indeed, it is generally recognised that
methanol forms from CO2, while CO acts as scavenger of oxygen atoms coming from
water production, hindering active metal sites [2, 4, 17, 20–22]. Then, water forma-
tion explains a lower functionality in CO2-rich syngas streams of traditional alumina-
based synthesis catalysts in comparison to zirconia ones [16, 23–28]. In fact, though
the enhanced reactivity of ZrO2-promoted Cu systems is generally ascribed to a lower
water affinity than commercial alumina-based ones, valuable promoting effects of
many oxide promoters and carriers on the functionality of Cu catalyst have been
documented [2, 4, 28–32]. In this context, the lack of definitive evidences on the CO2-
hydrogenation mechanism of oxide-promoted Cu systems represents to date a major
drawback for further catalyst advances [2, 29–32]. Despite the methanol synthesis
reactions network on metal Cu has been thoroughly described by a classical
Langmuir–Hinshelwood (L–H) model [17, 33, 34], a mix of structural, electronic
and chemical effects is invoked to explain the positive influence of oxide promoters
on the CO2-hydrogenation pattern of Cu catalyst [2, 4, 28–32].
In fact, apart from the generally recognised indispensable promoting influence of
ZnO on the catalytic functionality of Cu [2, 4, 30, 33, 35–39], the lack of relation-
ships between metal exposure (MSA) and activity seems to depend on a strong
“synergism” of metal and oxide phases controlling texture, exposure of active sites
and interaction with reagents, products and intermediates.

5.2 Thermodynamic Aspects of Methanol Synthesis


Reactions

Since almost a century, methanol synthesis is one of the most important processes
of the chemical industry, methanol, with ammonia and sulphuric acid, being one of
the three most produced chemicals in the world. With an overall capacity of about
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 105

Table 5.1 List of reactions of the methanol synthesis processes


Description Reaction stoichiometry ΔH0 (kJ/mol) Acronym
Methanol synthesis CO2 + 3 H2 ⇆ CH3OH + H2O 49.4 MS
Reverse water gas shift CO2 + H2 ⇆ CO + H2O +41.1 RWGS
CO hydrogenation CO + 2 H2 ⇆ CH3OH 90.5 COH
Methanol steam reforming CH3OH + H2O ⇆ CO2 + 3 H2 +49.4 MSR
Water gas shift CO + H2O ⇆ CO2 + H2 41.1 WGS
Methanol decomposition CH3OH ⇆ CO + 2 H2 +90.5 MD
ΔH0 at standard conditions (25  C and 0.1 MPa)

60 Mton per year [40], nowadays methanol synthesis is run on Cu–ZnO/Al2O3


catalysts in the range of 220–300  C, at a pressure of 5–10 MPa, with a syngas feed
generally containing 10–20 % of COx in H2 [2–6, 21, 40–45].
Indeed, whatever feed composition, the chemistry of methanol synthesis processes
can be described by two main reactions, both involving CO2 as reactant: the methanol
synthesis (MS) and the reverse water gas shift (RWGS). A linear combination of MS
and RWGS accounts for the straight CO hydrogenation to methanol (COH), besides
all those relevant processes involving CO, CO2, H2, CH3OH and H2O, like methanol
steam reforming (MSR) and decomposition (MD) and water gas shift (WGS),
constituting the reverse of COH, MS and RWGS, respectively (Table 5.1).
Due to the involvement of the same molecules, and perhaps of similar intermedi-
ates, the above reactions deserve analogous catalyst formulations, including copper
as the active phase [1–6, 21, 33, 40–45]. In addition, they occur in a relatively narrow
range of temperature (200–300  C), the lower limit relying on kinetic factors, while,
at industrial scale, yield optimisation deserves different operating pressures.
The optimum operating conditions of each process are, in fact, evident from
Fig. 5.1a, showing an overview of the free energy of MS, RWGS and COH reactions
in the range of 25–300  C. Both MS and RWGS reactions are characterised by ΔG
values that are always positive, rising for the former (4!56 kJ/mol) and decreasing
for the latter (28!17 kJ/mol), while COH is the only favoured process (ΔG < 0)
below 150  C. Furthermore, the MS and COH reactions are exothermic and accom-
panied by a decrease in volume, while the endothermic RWGS is favoured at higher
temperatures and insensitive to pressure (Table 5.1).
Therefore, whether CO or CO2 is the C-atom source, thermodynamic data
indicate that methanol formation is favoured at lower temperatures and higher
pressures as shown, in particular, by the equilibrium conversion-selectivity data
of a stoichiometric (1:3) CO2:H2 mixture in the range of 160–260  C (Fig. 5.1b).
At any pressure and below 160  C, thermodynamic data show full selectivity to
methanol, decreasing with temperature more steeply as pressure decreases. The
CO2 conversion lowers with temperature, though its variation is very sensitive to
pressure at 160  C (19–43 %) and almost negligible (20–25 %) at 260  C. This is
because of the prevailing contribution of MS over the RWGS reaction at lower
temperature and vice versa (Fig. 5.1a).
Finally, thermodynamic data deserve a final remark concerning the methanol
C-atom source. In fact, CO in the syngas feed for the methanol synthesis does not
106 F. Arena et al.

a 60 b 1.0
XCO2
6 MPa SCH3OH
50 SCO
4 MPa
0.8 2 MPa
40

30 RWGS 1 MPa
ΔGreaz (kJ/mol)

0.6

ΧCO2, SX
20
1 MPa
6 MPa 4 MPa
10 MS
0.4
2 MPa
4 MPa
0
2 MPa
-10 1 MPa
0.2
COH
-20
6 MPa
-30 0.0
25 75 125 175 225 275 160 180 200 220 240 260
Temperature (°C) Temperature (°C)

Fig. 5.1 Free energy of the MS, RWGS and COH reactions (a) and equilibrium conversion-
selectivity values of the CO2-hydrogenation reaction at various pressures (b)

take part directly in the reaction, but it rather does influence the process through two
effects: (i) producing the “reactive” CO2 intermediate via the WGS reaction and
(ii) influencing morphology, state and composition of metal Cu particles. Thermo-
dynamic data, however, suggest that it would be advantageous to the synthesis of
methanol by direct hydrogenation of CO, the COH reaction being characterised by
lower ΔG values in the concerned range of temperature (Fig. 5.1a). At any pressure,
it would allow much higher conversion and yield per pass than current processes,
proceeding through CO2 hydrogenation, as demonstrated by the CO and CO2-
hydrogenation equilibrium values reported in Table 5.2.
Therefore, the discovery of novel catalysts producing methanol with high
selectivity, directly by hydrogenation of CO (methanation is the competitive
process), would represent a further decisive breakthrough for the MOH synthesis
technology [33].

5.3 Catalysts for Methanol Synthesis

Although the thermodynamic analysis indicates low temperature and high pressure
as the optimal reaction conditions for the CO2-hydrogenation process, the activa-
tion of carbon dioxide occurs with a sufficient rate only at temperatures higher than
200  C. At such temperatures, however, the formation of by-products like CO,
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 107

Table 5.2 Equilibrium conversion values of stoichiometric CO–H2 (1:2) and CO2–H2 (1:3)
mixtures at various temperatures and pressures

Temperature Pressure
2 MPa 4 MPa 6 MPa 2 MPa 4 MPa 6 MPa
 a b
( C) XCO (%) XCO2–SCH3OH–YCH3OH (%)
180 66 77 82 24-92-22 35-97-34 40-98-40
200 53 68 75 20-80-16 29-93-27 35-96-34
220 39 57 65 19-59-11 25-83-22 32-85-27
240 26 44 55 19-36-7 23-66-15 28-79-22
260 15 32 43 21-8-2 22-46-10 25-66-17
a
Assuming a total selectivity to methanol (i.e. XCO ¼ YCH3OH)
b
From MS and RWGS equilibrium data (Fig. 5.1b)

hydrocarbons and alcohols lowers methanol selectivity and yield; then, preferably,
catalysts should be able to activate the CO2 at temperatures as low as possible,
avoiding undesired by-products formation.
Several surveys have been devoted to catalyst design, investigating the role of
different active phases and their interaction with many oxide support and promoters
[2, 4].

5.3.1 Active Phases

Although a variety of metal phases were investigated, a competitive catalyst


formulation for the CO2-hydrogenation process remains that of conventional synthe-
sis catalysts, including copper and zinc oxide on an alumina carrier, in different
proportions [2, 4, 6, 41]. Generally, CuO loading is comprised between 40 and 80 %,
ZnO between 10 and 30 %, while Al2O3 represents the remaining weight fraction,
being comprised between 10 and 20 % [6]. These require extremely pure syngas
(i.e. S- and Cl-compound concentrations lower than 0.1 ppm) and mild operating
conditions to avoid the premature ageing caused by the copper sintering [2, 41].
Then, relatively large amounts of water produced by MS and RWGS (Table 5.1)
inhibit catalytically active sites, probably constituting the hardest drawback of
conventional methanol synthesis catalysts, accelerating the catalyst sintering and
the synthesis rate decay [4, 22–25, 45, 46]. Indeed, the presence of CO in syngas,
besides serving as CO2 source, acts as scavenger of oxygen atoms, hindering the
negative effects of water formation [4, 20, 33].
Hence, catalysts for carbon dioxide hydrogenation should exhibit a higher water
tolerance, suggesting that this could be at the origin of a lower functionality of the
alumina-based catalysts in comparison to zirconia ones [4, 16, 24–30].
In combination with different promoters, copper is to date considered as the best
active phase for the methanol synthesis reactions [2–4, 6, 31, 33, 47], as
documented by a brief summary of activity data of the most used metal catalysts
in Table 5.3. In particular, Cu-based catalysts exhibit the best CO2-hydrogenation
108 F. Arena et al.

Table 5.3 Activity data in the CO2-hydrogenation reaction of different catalysts


Methanol
Catalyst Preparation method P(MPa) T ( C) CO2 conv. (%) yield (%) Ref.
Cu–ZnO/ZrO2 Co-precipitation 8 220 21.0 14.3 [49]
Ag/ZnO/ZrO2 Co-precipitation 8 220 2.0 1.9 [49]
Au/ZnO/ZrO2 Co-precipitation 8 220 2.5 1.5 [49]
Pd–CeO2 Impregnation 2 250 4.1 1.2 [50]
Pd/Zn/CNTs Incipient wetness 3 250 6.3 6.3 [51]
G2O3–Pd/SiO2 Incipient wetness 3 250 n/a – [52]
LaCr0.5Cu0.5O3 Sol-gel 2 250 10.4 9.4 [53]
Pt/CeO2 Impregnation 30 230 8.1 5.5 [54]
Ru/CeO2 Impregnation 30 230 8.2 0 [54]
Ni/CeO2 Impregnation 30 230 2.5 0 [54]

functionality to methanol, while, though active, Ni, Ru and Rh mainly encourage


the formation of methane [47, 48]. Examples of less reactive metal components are
Pd, Pt, Ag and Au, also catalysing the formation of high amounts of carbon
monoxide and methane [49–54]. Moreover, Pd–ZnO catalyst supported on
multiwalled carbon nanotubes (CNTs) offers very high methanol selectivity at
very low conversion per pass, while a Pd/CeO2 system forms preferentially form-
aldehyde instead of methanol (Table 5.3).
Despite the unpredictability of the catalyst performance at microscopic level,
such evidences indicate that the superior methanol synthesis functionality of Cu is
likely associated to its peculiar ability to hinder the breaking of the C–O bond, at
variance of more active hydrogenation catalysts, like Ni, Pt, Ru and Pd, forming
huge amounts of methane.
Furthermore, a very controversial issue concerns the chemical state of the active
sites under stationary reaction conditions [2, 4, 11, 14–18, 33, 55]. Indeed, some
claim that metal Cu atoms are uniformly active, according to a catalytic activity
comparing the metal surface area [35–58], whereas others report that electron-
deficient Cuδ+ species dissolved across the ZnO promoter are the active sites, on the
basis of a superior specific activity than un-promoted systems [4, 12, 59–69]. Any-
how, strong synergistic effects on the functionality of the binary Cu–ZnO system
can be grouped into the following six categories [70]:
– Formation of Cu+ ions on the ZnO phase [59, 60, 71]
– Schottky-type junction effects at the Cu–ZnO interface [61]
– Electronic interactions between Cu and ZnO [61]
– Formation of Cu–Zn pair alloy [35, 39]
– Specific reaction at the Cu–ZnO interface [65]
– Stabilisation of Cu in a morphologically active form by ZnO [69]
Metal copper centres are by far the most accepted species responsible for the
catalytic activity, though their exact nature is not fully clear, strongly depending on
the gas phase composition.
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 109

Fig. 5.2 Schematic model


for the wetting/non-wetting
phenomena of the Cu
particles on the ZnO support
[70]: (I ) round-shaped
particle under oxidising
syngas conditions; (II)
disc-like particle under
more reducing conditions;
(III) surface Zn–Cu alloying
due to stronger reducing
conditions; (IV) brass alloy
formation due to severe
reducing conditions

For instance, on the basis of DFT calculations, several authors conclude that
open Cu surface partially covered by oxygen provides a better model for the active
sites of methanol synthesis [17, 72, 73]. On the other hand, higher reducing
potentials increase oxygen vacancies at the ZnO surface and spread over Cu
crystallites [33, 64], ultimately resulting in the formation of Cu–Zn surface alloys
[65, 66]. Reversible changes in the Cu coordination number with the particle shape
on the ZnO support induced by the wetting/non-wetting phenomenon were also
reported [64]. Under oxidising conditions, the ZnO surface is not wetted by copper,
leading to more spherical-shaped copper particles, while a wetting transition takes
place if the ZnO surface becomes oxygen deficient as the copper particles assume a
“pancake” morphology, as shown in Fig. 5.2 for the wetting/non-wetting model of
Cu particles on the ZnO support [64]. HRTEM investigations show that the changes
for ZnO-supported samples are caused both by adsorbate-induced changes in
surface energies and changes in the interfacial energy [33, 74].
Therefore, the observed dynamic restructuring in response to changes in the
properties of the reaction environment demonstrates that the active sites for meth-
anol catalysts are essentially generated during reaction [17, 33, 72–74].
110 F. Arena et al.

5.3.2 Supports and Promoters

Although the superior reactivity of ZnO-promoted Cu catalysts is generally


recognised, valuable promoting effects of many oxide promoters and carriers on
the functionality of Cu catalyst have been documented.
Therefore, novel catalyst formulations and preparation methods have been
proposed to improve the process performance in terms of activity, selectivity and
stability [2, 4].
Among the oxide promoters of the Cu–ZnO system, ZrO2 is of overwhelming
interest because of its good mechanical and thermal stability under reducing and
oxidising atmospheres, high specific surface area and a peculiar surface function-
ality [16, 25, 28, 70, 75]. It is not a case that pure ZrO2 shows distinct activity and
selectivity for methanol synthesis at higher reaction temperatures [70].
Then, adopting a new preparation method based on the reverse co-precipitation
of precursors under ultrasound irradiation, Arena et al. synthesised Cu–ZnO/ZrO2
catalysts with surface exposure up to 170 m2/g and metal dispersion and surface
area values up to 60 % and 60 mCu2gcat1, respectively [25]. Recently, Guo
et al. obtained Cu–ZnO/ZrO2 catalysts with favourable characteristics such as
small grain size and low reduction temperature via the urea–nitrate combustion
method [76]. The presence of urea distributes the combustion heat, allowing a very
fast quenching effect leading to very small CuO particles that favour a stronger
interaction between copper, ZnO and ZrO2. An increase in the urea content led to
the partial transformation of t-ZrO2 to m-ZrO2, stabilising higher concentration of
the reactive intermediates that improve methanol selectivity [76].
Słoczyński et al. studied the effects of Ga, B, In, Gd, Y, Mg and Mn on Cu–ZnO/
ZrO2 catalyst, pointing out a remarkable influence on the catalyst activity [32]. The
highest methanol yield was obtained by the addition of Ga, while In implies a
severe activity loss. Toyir et al. reported that the addition of Ga2O3 to the Cu–ZnO
system enhances the methanol production, due to selectivity values two times
higher than that of the corresponding Cu–ZnO system [77]. Finally, studying the
influence of Ce addition on Cu/CexZr(1-x)O2 catalyst, Pokrovski and Bell observed a
strong increase of the methanol selectivity [31], later confirmed by Arena et al.,
though a low surface exposure depresses the catalyst productivity [28–31].

5.4 Mechanistic–Kinetic Aspects of CO2-Hydrogenation


Reactions

5.4.1 Methanol Synthesis

Although bare metal Cu exhibits its own reactivity in the hydrogenation of CO2 to
methanol, well described by macro- and microkinetic models [17, 33, 72, 73], the
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 111

promoting effect of many oxide systems reflects the occurrence of structural,


chemical and electronic effects, enhancing exposure and reactivity of Cu sites. In
addition, it must be stressed that oxide promoters feature a peculiar surface affinity
to either CO2 or H2 molecule that can further contribute to shape the functionality of
multicomponent catalytic systems [28, 29, 36, 75–79].
Although this partially explains the high uncertainty on active sites in charge of
the various surface functionalities of promoted Cu catalysts [2–4] and despite the
similar physico-chemical characteristics of catalysts for methanol synthesis (MS),
water gas shift (WGS), methanol steam reforming (MSR) and methanol decompo-
sition (MD) processes [2, 4, 43, 44, 80–84] and composition and redox potential of
the reacting atmosphere also influence morphology and chemical state of Cu
particles indeed [85, 86].
In this respect, Klier et al. propose that oxidised Cu is active and reduced Cu is
not in methanol synthesis, arguing that copper incorporated into Zn lattice, as Cu+,
is the active site for reactant adsorption [35, 83].
Quanto-chemical calculations lead several authors to conclude that MS
and RWGS might not occur on clean Cu <111> surface [17, 72, 73]. However,
Yang et al. suggest that methanol formation from both CO2 and CO, along with
the WGS–RWGS interconversion paths, occurs via the carboxyl (HOC*O) route,
in agreement with the conclusions of Zhao et al., claiming yet a fundamental
influence of the surface Cu oxygen coverage on the hydrocarboxyl route [72, 73].
In contrast, the microkinetic model developed by Grabow and Mavrikakis supports
the formate reaction path, still driven by partially oxidised Cu facet (CO2 ! HCOO*
! HCOOH* ! CH3O2* ! CH2O* ! CH3O* ! CH3OH*) [17]. Both Cu+ and
Cu0 seem to be active in the MD reaction, though the reducing environment
suggested that the latter are the active sites [84], while Choi and Stenger came to
conclusion that Cu2+ centres drive both MD and MSR reactions [44].
This “mimetic” nature of active sites depends on the fact that, despite all
the above reactions involve the same molecular species, formation and fate of the
numerous reaction intermediates strictly depend on the state of the catalyst surface
under stationary conditions. For instance, while MSR and WGS reactions imply
mostly oxidative conditions, MS, RWGS and MD proceed in a reducing environ-
ment. In other words, each process determines a different state of metal sites,
shaping the reactivity of the active Cu phase in MS, WGS–RWGS, MD and MSR
processes.
In this context, it is generally recognised as the systematic promoting influence
of ZnO on the reactivity of Cu, because of a peculiar solid-state interaction pattern
leading to the effects of different nature. In particular, a recognised strong Cu–ZnO
synergism enhancing the catalytic functionality of the active phase has been mostly
explained by four types of physical and chemical effects [2, 4, 33, 39, 71]:
(i) ZnO enhances metal dispersion, exerting a structural effect on Cu phase.
(ii) ZnO acts as a reservoir for atomic hydrogen that, spilling over the Cu surface,
speeds up the hydrogenation of the formate intermediate.
112 F. Arena et al.

(iii) ZnO confers a peculiar morphology to Cu particles, by stabilising specific


active surface planes.
(iv) ZnO creates additional active sites on the Cu surface.
In addition, a peculiar solid-state interaction pattern renders the Cu–ZnO system
very dynamic, accounting for marked changes in the morphology of metal particles,
associated with the extent that the metal particles wet the underlying support, with
composition and reduction potential of the gas mixture [64, 85, 86]. The dynamics
of particle morphology can also be used to counteract the effect of sintering of
copper particles [33], though morphology changes predicted by the Wülff recon-
struction model were questioned [86]. Then, while a strong Cu–ZnO synergism
positively influences metal dispersion, the nature of the promoting effect of ZnO is
still controversial. Further possible electronic/chemical effects (ii–iv) due to the
intimate contact of metal/oxide phases cannot, yet, be excluded [72, 73].
The picture is further complicated by the fact that, besides Cu and ZnO precursors,
catalyst formulations generally involve an additional (structural) promoter, improving
the dispersion and the resistance to sintering of Cu and ZnO phases [2, 4, 6]. Due to a
peculiar capability to interact with both Cu and Zn precursors, stabilising highly
dispersed, prevalently amorphous phases, ZrO2 and Al2O3 are to date the most
effective “carriers” of Cu–ZnO systems. Particularly, ZrO2-based systems feature a
superior activity than conventional Al2O3-based synthesis catalysts, probably because
of a superior hydrophobicity favouring water desorption [2, 4, 25–28, 76, 87]. Never-
theless, as documented by Fischer and Bell, zirconia exhibits also a peculiar surface
affinity to CO2 playing an “active” role on the reactivity pattern of the Cu/ZrO2
catalyst, favouring the formation of the reactive formate intermediate [75].
This seems not a peculiarity, since the formation of various reaction intermediates
and precursors at the surface of Al2O3, SiO2 and TiO2 carriers was also reported [88].
Then, a network of physical and chemical effects, due to the strong synergism of
the various phases, shapes the catalytic pattern of multicomponent systems, dis-
guising the role and influence of each single promoter on the structure and reactivity
of the active phase.
Taken as physical location of active sites, mostly as Cu+ centres, the metal/oxide
interface is in fact generally considered as a measure of the Cu–oxide(s) “synergism”
[11, 26, 27, 35, 63]. Its crucial influence on the catalytic functionality could explain
the lack of direct relationships between activity and metal surface area, recorded for
both Cu–ZnO/Al2O3 [87] and Cu–ZnO/ZrO2 systems [25–30].
This denotes a variable specific activity of Cu sites (e.g. TOF) with
metal dispersion (Fig. 5.3), supporting the hypothesis that the reactivity of
multicomponent catalysts no longer depends only on the availability of Cu sites
[11, 25–30, 87, 89]. Particularly, a peculiar surface affinity to CO2, enhancing the
formation of reactive intermediates, as a key factor of the reactivity of promoted Cu
catalysts could be taken into account [28, 29, 36, 87–90].
Due to the unfeasibility of getting direct information on the extent of the metal/
oxide interface, our first attempts to highlight the key factor(s) controlling the
reactivity of multicomponent Cu–ZnO catalysts were based on the evaluation of
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 113

a 0.030 b 0.030
TOFCO2 TOFCO2
TOFCH3OH TOFCH3OH
0.025 0.025

0.020 0.020

TOF (s-1)
TOF (s-1)

0.015 0.015

0.010 0.010

0.005 0.005

0.000 0.000
0 10 20 30 40 50 60 0 10 20 30 40
D (%) dCu (nm)

Fig. 5.3 Influence of the metal dispersion (a) and average metal particle diameter (b) on the
turnover frequency of CO2 conversion and CH3OH formation (Reprinted from Ref. [25]. Copyright
2007, with permission from Elsevier)

the “θ” factor (θ ¼ [4(MSAOSA)/(SA)2], MSA, OSA and SA being metal, oxide
and total surface area, respectively), taken as a measure of the degree of contact of
metal and oxide phases [11].
Experimental data show typical “volcano” trends (Fig. 5.4), supporting the
thought that CO2-hydrogenation systems might bear a balanced exposure of metal
(MSA) and oxide (OSA) surface [11]. This also explains the negative effect of an
overwhelming MSA exposure on the reactivity of Cu–ZnO and Cu–ZnO/Al2O3
catalysts [87].
Chemisorption and FTIR studies on high surface area Cu–ZnO/ZrO2 catalysts
[26, 27] prove that the interaction of Cu particles with ZnO and ZrO2 phases leads to
stabilisation of Cuδ+ sites at the metal/oxides interface, to an extent (i.e. Cuδ+/Cu0)
comparable with the CO/N2O uptake ratio [26].
Then, the normalisation of TOF to the CO/N2O uptake (TOF1) and oxide to
metal surface area (OSA/MSA) ratios (TOF2) results in two straight-line relation-
ships, fairly insensitive to Cu dispersion (Fig. 5.5), supporting a direct influence of
the oxide phases on the functionality of promoted Cu catalysts.
This hypothesis is supported by the conclusions of a comparative study on the
effects of Ga2O3, Cr2O3, Al2O3 and ZrO2 carriers on the reactivity of the Cu–ZnO
system, showing different effects depending on the nature of oxide promoter [77].
In particular, these were classified into two basic categories:
• Al2O3 and ZrO2 improve the metal dispersion, acting essentially as structural
promoters of the active Cu phase.
114 F. Arena et al.

Fig. 5.4 Relationships 11e-3


between TOF and θ at
180 (Δ), 200 (☐) and (○) 10e-3
220  C (Reprinted from
Ref. [11]. Copyright 2004, 9e-3
with permission from
Elsevier) 8e-3

7e-3

TOFCH3OH (s-1)
6e-3

5e-3

4e-3

3e-3

2e-3

1e-3

0e-3
0.0 0.2 0.4 0.6 0.8 1.0
q

101

100
TOF1
10-1
TOF⬘s (s-1)

TOF
10-2

TOF2

10-3 0 10 20 30 40 50 60
D (%)

Fig. 5.5 Influence of metal dispersion on the turnover frequency (TOF) and TOF normalised
to the CO/N2O (TOF1) and OSA/MSA (TOF2) ratios, respectively (Reprinted from Ref.
[26]. Copyright 2008, with permission from Elsevier)
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 115

• Ga2O3 and Cr2O3 enhance the specific activity of surface metal Cu atoms, acting
thus as chemical/electronic promoters of the active Cu phase [77].
Likewise, the addition of ceria and ceria-yttria yields a remarkable improvement
in the reactivity of the Cu/Al2O3 system, which was attributed to the generation of
surface oxygen vacancies at ceria surface favouring CO2 activation [89].
The peculiar promoting effect of ceria was confirmed by several studies on the
influence of either ZnO or ZrO2 substitution in Cu–ZnO/ZrO2 catalysts
[28–31]. Apart from a general negative influence of either ZnO or ZrO2 replacement
on total and metal surface area [28–30], activity data substantially confirm the
requirement of the ZnO precursor and a progressive enhancement in the surface
reactivity of the Cu–ZnO system with the extent of CeO2–ZrO2 substitution [30, 31].
Due to a much lower efficiency as structural promoter than Al2O3 and ZrO2, resulting
in considerably lower SA and MSA exposure, however, the Cu–ZnO/CeO2 system
features a lower productivity in comparison to Cu–ZnO/Al2O3 and Cu–ZnO/ZrO2
catalysts [28, 29].
Nevertheless, the lack of direct relationships between activity and MSA from the
specific metal surface activity data of the Cu–ZnO/Al2O3, Cu–ZnO/ZrO2 and
Cu–ZnO/CeO2 catalysts is also evident, in this case, documenting the superior func-
tionality of the ceria-supported catalyst at any temperature and pressure (Fig. 5.6).
Chemisorption data of the various catalysts at “steady-state” conditions, probed
by temperature-programmed measurements, show a complex pattern of CO2 and H2
desorption (Fig. 5.7a), similar for each catalyst, along with a band of CO formation,
likely due to an incipient reduction of CO2 [28, 29]. These account for an extensive
surface COx coverage that, referred to metal Cu surface area, gives a decreasing
trend with dispersion, as shown in Fig. 5.7b. The variations are less pronounced per
unit of total (SA) and oxide surface area (OSA). The variations are less pronounced
per unit of total (SA) and oxide surface area (OSA).
An extent of conversion of 60–80 % to MOH and CO under H2 stream disclosed
the reactive nature of the adsorbed CO2 species [29], while the metal-specific surface
activity of the Cu–ZnO/Al2O3, Cu–ZnO/ZrO2 and Cu–ZnO/CeO2 catalysts depicts
decreasing trends with dispersion (Fig. 5.8), similar to that found for Cu–Zn/ZrO2
catalysts (Fig. 5.3) and comparable with the COx uptake in Fig. 5.7b.
On the other hand, unchanging values of the activity data normalised to the oxide-
specific surface area (Fig. 5.8) suggest again a fundamental contribution of the oxide
carrier on the CO2-hydrogenation functionality of the Cu–ZnO system [28, 29].
Therefore, steady-state chemisorption data and structure–activity relationships
suggest that oxide promoters directly affect the catalytic functionality of metal Cu,
likely determining an additional availability of CO2 adsorption–activation sites,
enhancing the surface concentration of the reactive intermediate(s) [29]. On this
account, the superior performance of the Cu–ZnO/ZrO2 catalyst might mirror a
crucial influence of the zirconia carrier on the availability and reactivity of surface
sites [25–30].
Finally, experimental evidences concerning the reaction network leading to MOH
and CO formation generally show a positive effect of the space velocity on methanol
116 F. Arena et al.

0.8
CuZnAl 3.0 MPa
CuZnZr
SMSA (mmolCO2 m Cu-2 h-1) CuZnCe
0.6

0.4

0.2

0.0
453 473 493 513
Temperature (K)

0.8
CuZnAl
CuZnZr
SMSA (mmolCO2 m Cu-2h-1)

CuZnCe
0.6

0.4

0.2

5.0 MPa
0.0
453 473 493 513
Temperature (K)

Fig. 5.6 Specific metal surface rate of CO2 conversion in the range of 453–513 K (Adapted from
Ref. [29]. Copyright 2013, with permission from Elsevier)

selectivity [28–30, 87], resulting in steadily decreasing selectivity-conversion rela-


tionships [25–27]. Altogether these evidences indicate the occurrence of a consecu-
tive reaction network leading to the primary formation of methanol (MS) at low
temperature (e.g. T < 200  C) and high pressure [28, 29]. The formation of CO
occurs mostly at higher temperature (T > 200  C) via the consecutive and parallel
MD and RWGS endothermic paths, respectively [25, 28, 29, 44, 87, 91].
Then, the mechanistic scheme proposed in Fig. 5.9 emphasises the cooperative
role of the various phases on the main reaction pathway driving the primary
formation of methanol at the metal/oxide interface [28].
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 117

a 1×10-10 b 30
H2

1×10-10 500K MSA

Surface COx release (μmol⋅m-2 )


8×10-11
desorption rate (a. u.)

20
6×10-11
443K
CO
5×10-11

466K 525K
4×10-11 CO2 OSA
10

3×10-11
CuZnAl
CuZnCe SA
2×10-11 CuZnZr

1×10-11 0
293 333 373 413 453 493 533 573 5 10 15 20 25 30 35
Temperature (K) D (%)

Fig. 5.7 (a) Temperature-programmed desorption data of CO2, CO and H2 of CuZnAl, CuZnZr
and CuZnCe catalysts after treatment under “steady-state” reaction conditions; (b) relationships
between the “steady-state” adsorption capacity referred to total (SA), metal (MSA) and oxide
surface area (OSA) and metal dispersion (Adapted from ref. [28, 29], Copyright 2013, with
permission from Elsevier)

0.4
Legend
Specific surface rate (μmolCO2⋅m-2⋅s-1)

MSSA OSSA
0.3
3.0 MPa
5.0 MPa

0.2

0.1

0.0
0 10 20 30
D (%)

Fig. 5.8 Metal (MSSA)- and oxide (OSSA)-specific surface rates of CO2 conversion vs. dispersion
(Adapted from Ref. [28]. Copyright 2013, with permission from Elsevier
118 F. Arena et al.

Fig. 5.9 Surface reaction mechanism of CO2 hydrogenation to methanol on supported Cu–ZnO
catalysts (Reprinted from Ref. [28]. Copyright 2013, with permission from Elsevier)

5.4.2 Direct Dimethylether Synthesis

The increasing demand of dimethylether (DME) as an alternative fuel to LPG and


gas oil, coupled to the possibility of realising an effective recycle of CO2 emissions,
explains the great scientific concern on the development of the novel “one-step”
DME synthesis processes via CO2 hydrogenation.
In fact, since methanol dehydration reaction occurs in a temperature window
compatible with that of the MS reaction, coupling the MS-RWGS and methanol
dehydration (MDH) reactions into a “one-step” process (2CO2 + 6H2 ⇆ C2H6O
+ 3H2O) provides a strong driving force for shifting the CO2 conversion above the
equilibrium values (2 CH3OH ⇆ C2H6O + H2O) forced by MS-RWGS reactions
(Fig. 5.1b), by in situ transformation of methanol to DME [2, 5].
Therefore, an effective synergism of MS-RWGS-MDH reactions results in remark-
able economic advantages of the overall CO2-hydrogenation process [2, 5]. However,
due to quite different catalyst requirements of synthesis and dehydration processes,
scouting attempts to develop bifunctional catalysts have been so far reported, using
various combinations of synthesis and dehydration catalyst formulations [92–94].
In particular, adopting a co-precipitating sedimentation method, Sun
et al. prepared multicomponent Cu–ZnO–ZrO2–Al2O3/HZSM5 catalysts further
promoted with a Pd/Al2O3 sample [92]. The synergistic effect of synthesis–dehy-
dration–hydrogenation functionalities results in a strong improvement of both CO2
conversion and MOH–DME selectivity which, at 200  C and 3.0 MPa, reaches the
remarkable level of 85–90 %, with an overwhelming formation of DME (75 %).
However, no definitive explanation to the strong promoting influence of Pd on the
functionality of the un-promoted system has been given [92].
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 119

Analogous results were obtained by Aguayo’s group using Cu–ZnO/Al2O3


synthesis catalyst in combination with either γ-Al2O3 or NaHZSM5 [93, 94].
Although having considerably lower DME selectivity values (T, 275  C; P,
4.0 MPa), a marked improvement of both conversion and selectivity led to overall
MOH–DME yields much higher than the single MOH synthesis process and above
the equilibrium values in the range of 240–340  C. They also provided a kinetic
model for the overall CO2 to DME hydrogenation process, coming to the conclusion
that the NaHZSM5 dehydration system could be more suitable for industrial purposes
as it offers a higher resistance to deactivation by water and coke formation and a
superior stability during the burn-off regeneration procedure [93].
Further attempts to synthesise DME via the CO2-hydrogenation process have been
made using “dual-bed” catalysts, obtained by physical mixing of synthesis and
dehydration catalysts. A strong improvement in the process performance of the
bifunctional Cu–ZnO/Al2O3/HZSM5 system was recently documented by Gao
et al. further to the addition of La [95]. In particular, a 2 wt% of La caused a fourfold
rise both in conversion and DME selectivity with respect to the un-promoted system,
ascribed to both an enhanced Cu dispersion and a superior strength of acid sites.
Although our preliminary attempt using Cu–ZnO/ZrO2 and HZSM5 systems did not
provide any significant improvement in CO2 conversion or oxygenated selectivity
[11], we addressed a more systematic investigation of the effects of the catalyst bed
configuration on the one-step DME synthesis performances of the hybrid Cu–ZnO/
ZrO2–HZSM5 system [96]. Among the various configurations, the “mono-bed”
configuration including pre-pelletised particles of both systems, with an optimum
weight ratio of 1, showed the highest level of synergism, with a marked improvement
of CO2 conversion, though at the expense of MOH–DME selectivity. This could
depend on the enhanced acidity or impurity present on the commercial HZSM5
system, somewhat promoting the MD reaction [96].
A preliminary study of the effects of the Ti/Zr ratio on the performances of
Cu–TiO2–ZrO2/HZSM5 bifunctional catalysts indicated a superior performance
further to the partial substitution (50/50) of zirconia with titania, though the effects
of the acidic catalysts were not addressed [97].
Dual-bed Cu–ZnO–Al2O3–ZrO2/HZSM5 configuration catalysts were also suc-
cessfully applied in the CO2-hydrogenation process, reaching conversion and
MOH–DME yield values of 0.31 and 0.21–0.06, respectively [98]. Based on such
experimental data, An et al. propose a CO2-hydrogenation process with CO recycle
to enhance both DME and MOH yields [98].

5.5 Synthesis Technologies

All commercial methanol technologies are made up of three main process sections
[6, 33, 41, 99–104]:
• Synthesis gas production
• Methanol synthesis
• Methanol purification
120 F. Arena et al.

In the design of a methanol plant, the three process sections may be considered
independently, and the technology may be selected and optimised separately for
each section. The criteria for the selection of technology are capital cost and plant
efficiency. The syngas preparation and compression typically account for
ca. 60 % of the investment, and almost all energy is consumed in this process
section.
The used synthesis gas is characterised by the so-called syngas module (M ),
defined by the ratio (H2–CO2)/(CO + CO2), where M ¼ 2 represents the stoichio-
metric synthesis gas for methanol formation [99, 100]. Hence, syngas with very
high CO/CO2 ratios reduces water formation and catalyst deactivation [6], while the
presence of inerts in the syngas composition reduces the methanol formation.
Typical inerts in the methanol synthesis are methane, argon and nitrogen.
Finally, the produced syngas needs a careful purification of the feedstock, since
the commercial methanol catalysts may be poisoned by sulphur, chlorine com-
pounds and metal carbonyls [6].
Currently syngas is obtained by three types of hydrocarbons reforming pro-
cesses: the steam reforming (SR), the partial oxidation (POX) and the auto-thermal
reforming (ATR).
The most commonly used method is the high energy-demanding SR process
carried out with excess steam-to-carbon ratios, which takes place in special alloy
tube reformers loaded with a nickel-based catalyst at 750–900  C. Burners
provide direct heat exchange. The exit syngas composition is a mixture of H2,
CO, CO2 and unreacted methane accounting for an M value of ca. 3. One of the
main drawbacks of this very complex technology is the long start-up time
increasing the operating costs.
Unlike the SR process, the POX technology presents a simple design with small
inert production without additional burners investments. It is a mature technology in
which natural gas or heavy hydrocarbons are mixed in a combustion process with
oxygen. The exothermic nature leads to a significant heat production requiring
external utilities (water and cold air). The main drawbacks are the comparatively
low hydrogen yield (M  1.8) and the tendency to produce soot at sub-stoichiometric
air/fuel ratios.
Finally, the ATR process carries out the natural gas reforming in a mixture of
steam and oxygen in the presence of a catalyst reducing the investment costs by
eliminating the expensive fired reformer technology. The produced heat is used to
supply the heat input needed by the steam reformer.
The synthesis gas produced presents an optimal module (2) for methanol
production, implying that it is deficient in hydrogen and contains high CO2 per-
centage [45]. The first ATR in operation in commercial scale at H2O/C ratio of 0.6
was an industrial demonstration in South Africa in 1999. Very large units were
started in South Africa in 2004 and Qatar in 2006 [101, 102]. A methanol plant with
a single line capacity of 10,000 MTPD is in the engineering phase for start-up in
Nigeria [103, 104].
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 121

5.5.1 Current Technologies

As evidenced from thermodynamic evaluations, MS and COH are exothermic


reactions. Then, low reaction temperature would combine high conversion with
low reaction rate. Using high amount of catalyst at lower temperature for better
selectivity or high temperature for higher productivity with high recycle ratio is a
possible strategy. Anyway, the optimal control of the reaction temperature is crucial
for facing the important drawback of methanol selectivity.
On this account, single bed converters, using the reactor as heat exchanger, and
multiple bed reactors, where the catalyst mass is separated in many sections by
cooling devices (heat exchangers or direct injections of cool gas), allow an optimal
temperature control.
Actually, the worldwide methanol production is essentially accomplished by the
ICI (ca. 60 % of world production), Lurgi (ca. 30 %) and Haldor Topsøe (ca. 10 %)
processes, including the following reactor technologies, all making use of the
Cu–ZnO/Al2O3 catalysts:
• Quench reactor
• Adiabatic reactors in series
• Boiling water reactors (BWR)
The ICI process (Fig. 5.10) uses an adiabatic reactor with four catalyst beds in
series (quench reactor), where the reaction is quenched by cold gas at several
points. Products are cooled by exchanging with the feed and water used for the
generation of high-pressure steam. Further cooling is accomplished by an
air-cooled heat exchanger in which methanol and water are condensed. The subse-
quent gas/liquid separation takes place in a flash drum under pressure, while the
unreacted gas is recycled after purging a small part, to keep the inerts level within
the operating limits. Finally the methanol purification is done in two separation
columns. In the first one, gases and other light impurities are removed, while in the
second, methanol is separated from other heavy alcohols.
The Lurgi process employs adiabatic reactors (Fig. 5.11), normally made up of
2–4 fixed beds, placed in series with cooling systems between the reactors
[105–107]. Cooling is achieved in several ways including preheating of high-
pressure boiler feed water, generation of medium-pressure steam and/or by preheat
of feed to the first reactor. This system features good economy of scale, and also the
mechanical simplicity contributes to lower investment costs. This technology,
called MegaMethanol, is developed for methanol capacities larger than one million
metric tons per year as that built in 2004 in Trinidad and Tobago.
Finally, the Haldor Topsøe methanol process (Fig. 5.12) makes use of the boiling
water reactors (BWR), constituted by a shell and tube heat exchanger with the
catalysts installed on the tube cooled by boiling water on the shell side [108]. By
controlling the pressure of the circulating boiling water, the temperature of the
reaction is controlled and optimised. This type of reactor is nearly isothermal,
122 F. Arena et al.

Fig. 5.10 Low-pressure methanol ICI process: (a) pure methanol column, (b) light ends column,
(c) heat exchanger, (d) cooler, (e) separator, (f) reactor, (g) compressor, (h) compressor recycle
stage [6]

Fig. 5.11 Scheme of the


Lurgi fixed-bed methanol
synthesis reactor (Reprinted
from Ref. [105]. Copyright
2012, with permission from
Elsevier)
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 123

Fig. 5.12 Scheme of the Haldor Topsøe boiling water reactor (BWR) (Reprinted from
Ref. [108]. Copyright 2009, with permission from IAHE)

giving a high conversion compared to the amount of catalyst installed therein.


Complex mechanical design of the BWR results in relatively high investment cost
and limits the maximum size of the reactor [108].

5.5.2 Is CO2 Hydrogenation to MOH–DME Feasible?

Since the current lack of real application instances, the feasibility of CO2 emissions
abatement is preliminarily assessed by a recent simulation study [109].
In particular, the study addresses the design of a CO2 to methanol process
hydrogenation, which uses flue gas of a coal thermoelectric plant [109]. Through
a detailed analysis of consumptions and releases, the study demonstrates that the
process feasibility depends on two essential requirements:
• The availability of hydrogen, produced, for instance, from electrolysis of water
by a zero-emissions electricity source (e.g. renewable and/or nuclear energy).
• The availability of waste heat in the power plant, to provide thermal energy to
the process. Indeed, in the absence of these thermal sources, the CO2 abatement
is close to 0.
An overview of the process is given in Fig. 5.13. Apart from the fundamental
relevance of the water electrolysis unit, the optimisation of which implies the
124 F. Arena et al.

Fig. 5.13 Block diagram of the CO2 abatement through the methanol synthesis process
(Reprinted from Ref. [109]. Copyright 2012, with permission from AIDIC)

exploitation of the oxygen generated by water splitting, the process includes a


chemical absorption CO2 capture unit from the flue gas of a power plant with
desulphurisation, using MEA as solvent. The synthesis reactor operates under
adiabatic conditions, and it is packed with a conventional Cu–ZnO/Al2O3 catalyst
for which, on the basis of available kinetic models, it is assumed a 33 % CO2
conversion to methanol and a recycle ratio of 5.0. On account of this, mass, energy
and CO2 balance analysis shows that it is possible to abate a large amount of CO2
from the production of methanol (up to 1.7 ton per ton of MOH) if “carbon-free”
hydrogen is available.
Furthermore, An et al. proposed a CO2 hydrogenation to DME hydrogenation
process [98]. Based on a kinetic model developed for the synthesis of DME from a
stoichiometric CO2–H2 mixture, they come to conclusion that 3 % of CO leads to a
strong increase in DME and MOH yields from CO2 hydrogenation, reaching at
270  C levels of 37 % and 6.5 %, respectively. Since CO is continuously produced
during the process, they show that its concentration can be tuned at a constant level
(3 %) in the reactor effluent recycle by proper optimisation of the process condi-
tions. A scheme of the CO2-to-DME process is shown in Fig. 5.14.
Despite the industrial feasibility of the synthesis of methanol from CO2 being
confirmed, only small demonstration units have been developed in recent years
worldwide. For instance, Carbon Recycling International (CRI) definitively proved
the feasibility of the carbon dioxide recycling with its first commercial scale plant in
Iceland, using available cheap geothermal energy sources to produce hydrogen by
water electrolysis [5]. Another pilot plant with an annual capacity of 100 ton is
under construction by the same company.
The CO2 hydrogenation has also been demonstrated on a pilot scale in Japan,
operating with a 50 kg CH3OH/day production and selectivity to methanol of
100 % [5].
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 125

Fig. 5.14 CO2 hydrogenation to DME process with CO recycle (Reprinted from Ref. [98].
Copyright 2008, with permission from American Chemical Society)

Further, exploring and developing new efficient catalysts formulations are


generating opportunities in concepts and technologies for the chemical industry.
Lurgi AG, for example, developed and thoroughly tested a high-activity catalyst for
methanol synthesis from CO2, obtaining an excellent selectivity at 260  C, while
the activity loss was comparable with that of commercial catalysts used in methanol
synthesis plants [5].

5.6 Conclusions

The hydrogenation to methanol and dimethylether might become a future prospective


of an efficient CO2 recycling, since, as effective substitutes of oil-derived fuels, they
would accomplish the strategic environmental goals of cutting greenhouse gas
emissions and reducing the air pollution levels in big metropolitan areas.
Although flue gas from fuel-burning power plants, cement and other factories
and syngas from ATR and POX processes are already suitable candidates for the
CO2-hydrogenation processes, however, a future scenario where the low concen-
tration atmospheric CO2 could be captured and recycled back to methanol/
dimethylether is predictable.
In this context, despite a century since its application, the catalysis of the
synthesis of methanol remains a tremendously fascinating topic, still worthy of
capturing a huge scientific concern worldwide. Hence, the fundamental research in
catalysis is destined to play a pivotal role in providing effective catalytic systems
operating at lower temperature and pressure that, in combination with novel
engineering approaches, are aimed at improving the feasibility of methanol/
dimethylether synthesis technologies.
126 F. Arena et al.

References

1. Our earth in 2050 (2013) http://greenphysicist2.blogspot.it. Accessed 28 Jun 2013


2. Wang W, Wang S, Ma X et al (2011) Recent advances in catalytic hydrogenation of carbon
dioxide. Chem Soc Rev 40:3703–3727
3. Song C (2006) Global challenges and strategies for control, conversion and utilization of CO2
for sustainable development involving energy, catalysis, adsorption and chemical processing.
Catal Today 115:2–32
4. Liu X-M, Lu G-Q, Yan Z-F et al (2003) Recent advances in catalysts for methanol synthesis
via hydrogenation of CO and CO2. Ind Eng Chem Res 42:6518–6530
5. Olah GA, Goeppert A, Surya Prakash GK (2009) Chemical recycling of carbon dioxide to
methanol and dimethyl ether: from greenhouse gas to renewable, environmentally carbon
neutral fuels and synthetic hydrocarbons. J Org Chem 74:487–498
6. Fiedler E, Grossmann G, Kersebohm DB, Weiss G, Witte C (2000) Methanol. Ullmann’s
encyclopedia of industrial chemistry. Wiley, New York
7. Larminie J, Dicks A (2003) Fuel cell systems explained, 2nd edn. Wiley, New York
8. Pontzen F, Liebner W, Gronemann V et al (2011) CO2-based methanol and DME efficient
technologies for industrial scale production. Catal Today 171:242–250
9. Naik SP, Ryu T, Bui V et al (2011) Synthesis of DME from CO2/H2 gas mixture. Chem Eng J
167:362–368
10. Ma J, Sun NN, Zhang XL et al (2009) A short review of catalysis for CO2 conversion. Catal
Today 148:221–231
11. Arena F, Spadaro L, Di Blasi O et al (2004) Integrated synthesis of dimethylether via CO2
hydrogenation. Stud Surf Sci Catal 147:385–390
12. Larson ED, Tingjin R (2003) Synthetic fuel production by indirect coal liquefaction. Energy
Sust Dev 7:79–102
13. Jia M, Li W, Xu H et al (2002) An integrated air-POM syngas/dimethyl ether process from
natural gas. Appl Catal A 233:7–12
14. Chinchen GC, Denny PJ, Jennings JR et al (1988) Synthesis of methanol: Part 1. Catalysts
and kinetics. Appl Catal 36:1–65
15. Bart JCJ, Sneeden RPA (1987) Copper-zinc oxide-alumina methanol catalysts revisited.
Catal Today 2:1–124
16. Yang C, Ma Z, Zhao N et al (2006) Methanol synthesis from CO2-rich syngas over a ZrO2
doped CuZnO catalyst. Catal Today 115:222–227
17. Grabow LC, Mavrikakis M (2011) Mechanism of methanol synthesis on Cu through CO2 and
CO hydrogenation. ACS Catal 1:365–384
18. Fujita S, Usui M, Ito H et al (1995) Mechanisms of methanol synthesis from carbon dioxide
and from carbon monoxide at atmospheric pressure over Cu/ZnO. J Catal 157:403–413
19. Sun Q, Liu Z (2011) Mechanism and kinetics for methanol synthesis from CO2/H2 over Cu
and Cu/oxide surfaces: recent investigations by first-principles-based simulation. Front Chem
China 6:164–172
20. Melian-Cabrera I, Granados ML, Fierro JLG (2002) Reverse topotactic transformation of a
Cu-Zn-Al catalyst during wet Pd Impregnation: relevance for the performance in methanol
synthesis from CO2/H2 mixtures. J Catal 210:273–284
21. Waugh KC (1992) Methanol synthesis. Catal Today 15:51–75
22. Sun Q, Liu C-W, Pan W et al (1998) In situ IR studies on the mechanism of methanol
synthesis over an ultrafine Cu/ZnO/Al2O3 catalyst. Appl Catal A 171:301–308
23. Inui T, Hara H, Takeguchi T et al (1997) Structure and function of Cu-based composite
catalysts for highly effective synthesis of methanol by hydrogenation of CO2 and CO. Catal
Today 36:25–32
24. Saito M, Fujitani T, Takeuchi M et al (1996) Development of copper/zinc oxide-based
multicomponent catalysts for methanol synthesis from carbon dioxide and hydrogen. Appl
Catal A 138:311–318
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 127

25. Arena F, Barbera K, Italiano G et al (2007) Synthesis, characterization and activity pattern of
Cu-ZnO/ZrO2 catalysts in the hydrogenation of carbon dioxide to methanol. J Catal
249:185–194
26. Arena F, Italiano G, Barbera K et al (2008) Solid-state interactions, adsorption sites and
functionality of Cu-ZnO/ZrO2 catalysts in the CO2 hydrogenation to CH3OH. Appl Catal A
350:16–23
27. Arena F, Italiano G, Barbera K et al (2009) Basic evidences for methanol-synthesis catalyst
design. Catal Today 143:80–85
28. Arena F, Mezzatesta G, Zafarana G et al (2013) Effects of oxide carriers on surface
functionality and process performance of the Cu-ZnO system in the synthesis of methanol
via CO2 hydrogenation. J Catal 300:141–151
29. Arena F, Mezzatesta G, Zafarana G et al (2013) How oxide carriers control the catalytic
functionality of the Cu-ZnO system in the hydrogenation of CO2 to methanol. Catal Today
210:39–46
30. Bonura G, Arena F, Mezzatesta G et al (2011) Role of the ceria promoter and carrier on the
functionality of Cu-based catalysts in the CO2-to-methanol hydrogenation reaction. Catal
Today 171:251–256
31. Pokrovski KA, Bell AT (2006) An investigation of the factors influencing the activity of
Cu/CexZr1-xO2 for methanol synthesis via CO hydrogenation. J Catal 241:276–286
32. Słoczyński J, Grabowski R, Olszewski R et al (2006) Effect of metal oxide additives on the
activity and stability of Cu/ZnO/ZrO2 catalysts in the synthesis of methanol from CO2 and
H2. Appl Catal A 310:127–137
33. Chorkendorff I, Niemantsverdriet JW (2005) Concepts of modern catalysis and kinetics.
Wiley, Weinheim
34. Askgaard TS, Norskov JK, Ovesen CV et al (1995) A kinetic model of methanol synthesis.
J Catal 156:229–242
35. Nakamura J, Uchijima T, Kanai Y et al (1996) The role of ZnO in Cu/ZnO methanol synthesis
catalysts. Catal Today 28:223–230
36. Ostrovskii VE (2002) Mechanisms of methanol synthesis from hydrogen and carbon oxides at
Cu-Zn-containing catalysts in the context of some fundamental problems of heterogeneous
catalysis. Catal Today 77:141–160
37. Granjean D, Pelipenko V, Batyrev ED et al (2011) Dynamic Cu/Zn interaction in SiO2
supported methanol synthesis catalysts unraveled by in situ XAFS. J Phys Chem C
115:20175–20191
38. Choi Y, Futagami K, Fujitani T et al (2001) The difference in the active sites for CO2 and CO
hydrogenations on Cu/ZnO-based methanol synthesis catalysts. Catal Lett 73:27–31
39. Fujitani T, Nakamura J (2000) The chemical modification seen in the Cu/ZnO methanol
synthesis catalysts. Appl Catal A 191:111–129
40. Global Methanol Market Review (2012) http://www.ptq.pemex.com. Accessed 24 Jun 2013
41. Hansen JB (1997) Methanol synthesis. In: Handbook of heterogeneous catalysis. Wiley,
Weinheim
42. Kakumoto T, Watanabe T (1997) A theoretical study for methanol synthesis by CO2
hydrogenation. Catal Today 36:39–44
43. Campbell CT, Daube KA (1987) A surface science investigation of the water-gas shift
reaction on Cu(111). J Catal 104:109–119
44. Choi Y, Stenger HG (2002) Fuel cell grade hydrogen from methanol on a commercial
Cu/ZnO/Al2O3 catalyst. Appl Catal B 38:259–269
45. Moulijn JA, Makkee M, van Diepen A (2001) Chemical process technology, 2nd edn. Wiley,
New York
46. Pipitone G, Bolland O (2009) Power generation with CO2 capture: technology for CO2
purification. Int J Grenh Gas Con 3:528–534
47. Wambach J, Baiker A, Wokaun A (1999) CO2 hydrogenation over metal/zirconia catalysts.
Phys Chem Chem Phys 1:5071–5080
128 F. Arena et al.

48. Razali NAM, Lee KT, Bhatia S et al (2012) Heterogeneous catalysts for production of
chemicals using carbon dioxide as raw materials: a review. Renew Sust Energ Rev
16:4951–4964
49. Słoczyński J, Grabowski R, Kozłowska A et al (2004) Catalytic activity of the M/(3ZnO ·
ZrO2) system (M ¼ Cu, Ag, Au) in the hydrogenation of CO2 to methanol. Appl Catal A
278:11–23
50. Shen W-J, Ichihashi Y, Ando H et al (2001) Effect of reduction temperature on structural
properties and CO/CO2 hydrogenation characteristics of a Pd-CeO2 catalyst. Appl Catal A
217:231–239
51. Liang X-L, Dong X, Lin G-D et al (2009) Carbon nanotube-supported Pd-ZnO catalyst for
hydrogenation of CO2 to methanol. Appl Catal B 88:315–322
52. Collins SE, Chiavassa DL, Bonivardi AL et al (2005) Hydrogen spillover in Ga2O3-Pd/SiO2
catalysts for methanol synthesis from CO2/H2. Catal Lett 103:83–88
53. Jia LJ, Gao J, Fang W et al (2009) Carbon dioxide hydrogenation to methanol over the
pre-reduced LaCr0.5O3 catalyst. Catal Commun 10:2000–2003
54. Tsubaki N, Fujimoto K (2003) Promotional SMSI effect on supported palladium catalysts for
methanol synthesis. Top Catal 22:325–335
55. Chinchen GC, Denny PJ, Parker DG et al (1987) Mechanism of methanol synthesis from
CO2/CO/H2 mixtures over copper/zinc oxide/alumina catalysts: use of 14C-labelled reac-
tants. Appl Catal 30:333–338
56. Deng J, Sun Q, Zhang Y et al (1996) A novel process for preparation of a Cu/ZnO/Al2O3
ultrafine catalyst for methanol synthesis from CO2 + H2: comparison of various preparation
methods. Appl Catal A 139:75–85
57. Pan WX, Cao R, Roberts DL et al (1988) Methanol synthesis activity of CuZnO catalysts.
J Catal 114:440–446
58. Ovesen CV, Stoltze P, Nørskov JK et al (1992) A kinetic model of the water gas shift
reaction. J Catal 134:445–468
59. Klier K (1982) Methanol synthesis. Adv Catal 31:243–313
60. Ponec V (1992) Cu and Pd, two catalysts for CH3OH synthesis: the similarities and the
differences. Surf Sci 272:111–117
61. Frost JC (1988) Junction effect interactions in methanol synthesis catalysts. Nature
334:577–579
62. Grunwaldt J-D, Molenbroek AM, Topsøe N-Y et al (2000) In situ investigations of structural
changes in Cu/ZnO catalysts. J Catal 194:452–460
63. Fujitani T, Nakamura J (1998) The effect of ZnO in methanol synthesis catalysts on Cu
dispersion and the specific activity. Catal Lett 56:119–124
64. Topsøe N-Y, Topsøe H (1999) FTIR studies of dynamic surface structural changes in
Cu-based methanol synthesis catalysts. J Mol Catal A 141:95–105
65. Greeley I, Gokhale AA, Kreuser J et al (2003) CO vibrational frequencies on methanol
synthesis catalysts: a DFT study. J Catal 213:63–72
66. Bower M, Hadden RA, Houghton H et al (1988) The mechanism of methanol synthesis on
copper/zinc oxide/alumina catalysts. J Catal 109:263–273
67. Yoshihara J, Campbell CT (1996) Methanol synthesis and reverse water-gas shift kinetics
over Cu(110) model catalysts: structural sensitivity. J Catal 161:776–782
68. Günter MM, Ressler T, Bems B et al (2001) Implication of the microstructure of binary
Cu/ZnO catalysts for their catalytic activity in methanol synthesis. Catal Lett 71:37–44
69. Bems B, Schur M, Dassenoy A et al (2003) Relations between synthesis and microstructural
properties of copper/zinc hydroxycarbonates. Chem Eur J 9:2039–2052
70. Vukojević S (2009) Copper colloids and nanoparticular Cu/ZrO2-based catalysts for metha-
nol synthesis. PhD Thesis, Ruhr-Universitat, Bochum
71. Fujitani T, Nakamura I, Ueno S et al (1997) Methanol synthesis by hydrogenation of CO2
over a Zn-deposited Cu(111): formate intermediate. Appl Surf Sci 121:583–586
5 Latest Advances in the Catalytic Hydrogenation of Carbon Dioxide. . . 129

72. Zhao Y-F, Yang Y, Mims C et al (2011) Insight into methanol synthesis from CO2
hydrogenation on Cu(111): complex reaction network and the effects of H2O. J Catal
281:199–211
73. Yang Y, Mims CA, Mei DH et al (2013) Mechanistic studies of methanol synthesis over Cu
from CO/CO2/H2/H2O mixture: the source of C in methanol and the role of water. J Catal
298:10–17
74. Hansen PL, Wagner JB, Helveg S et al (2002) Atom-resolved imaging of dynamic shape
changes in supported copper nanocrystals. Science 295:2053–2055
75. Fisher IA, Bell AT (1997) In-situ infrared study of methanol synthesis from H2/CO2 over
Cu/ZrO2/SiO2. J Catal 172:222–237
76. Guo X, Mao D, Lu G et al (2010) Glycine-nitrate combustion synthesis of CuO-ZnO-ZrO2
catalysts for methanol synthesis from CO2 hydrogenation. J Catal 271:178–185
77. Toyir J, Milouna R, Elkadri NE et al (2009) Sustainable process for the production of
methanol from CO2 and H2 using Cu/ZnO-based multicomponent catalysts. Phys Proceed
2:1075–1079
78. Fisher IA, Bell AT (1998) In situ infrared study of methanol synthesis from H2/CO over
Cu/SiO2 and Cu/ZrO2/SiO2. J Catal 178:153–173
79. Bianchi D, Chafik T, Khalfallah M et al (1994) Intermediate species on zirconia supported
methanol aerogel catalysts: III. Adsorption of carbon monoxide on copper containing solids.
Appl Catal A 112:57–73
80. Jones SD, Neal LM, Hagelin-Weaver HE (2008) Steam reforming of methanol using Cu-ZnO
catalysts supported on nanoparticle alumina. Appl Catal B 84:631–642
81. Mastalir H, Frank B, Szizybalski A et al (2005) Steam reforming of methanol over Cu/ZrO2/
CeO2 catalysts: a kinetic study. J Catal 230:464–475
82. Mei M, Xu L, Henkelman G (2009) Potential energy surface of methanol decomposition on
Cu(110). J Phys Chem C 113:4522–4537
83. Klier K, Chatikavanij V, Herman RG et al (1982) Catalytic synthesis of methanol from
CO-H2: IV. The effects of carbon dioxide. J Catal 74:343–360
84. Cheng W-H (1995) Reaction and XRD studies on Cu based methanol decomposition
catalysts: role of constituents and development of high-activity multicomponent catalysts.
Appl Catal A 130:13–30
85. Clausen BS, Schiøtz J, Gråbæk L et al (1994) Wetting/non wetting phenomena during
catalysis: evidence from in situ on-line EXAFS studies of Cu-based catalysts. Top Catal
1:367–376
86. Ressler T, Kniep BL, Kasatkin I et al (2005) The microstructure of copper zinc oxide
catalysts: bridging the materials gap. Angew Chem Int 44:4704–4707
87. Sun Q, Zhang Y-L, Chen H-Y et al (1997) A novel process for the preparation of Cu/ZnO and
Cu/ZnO/Al2O3 ultrafine catalyst: structure, surface properties, and activity for methanol
synthesis from CO2 + H2. J Catal 167:92–105
88. Bando KK, Sayama K, Kusama H et al (1997) In-situ FT_IR study on CO2 hydrogenation
catalysts supported on SiO2, Al2O3, and TiO2. Appl Catal A 165:391–409
89. Wang JB, Lee H-K, Huang T-J (2002) Synergistic catalysis of carbon dioxide hydrogenation
into methanol by yttria-doped ceria/γ-alumina-supported copper oxide catalyst: effect of
support and dopant. Catal Lett 83:79–86
90. Tang Q-L, Hong QJ, Liu Z-P (2009) CO2 fixation into methanol at Cu/ZrO2 interface from
first principles kinetic Monte Carlo. J Catal 263:114122
91. Yang Y, Evans J, Rodriguez JA et al (2010) Fundamental studies of methanol synthesis from
CO2 hydrogenation on Cu(111), Cu clusters, and Cu/ZnO(0001). Phys Chem Chem Phys
12:9909–9917
92. Sun K, Lu W, Wang M et al (2004) Low-temperature synthesis of DME from CO2/H2 over
Pd-modified CuO-ZnO-Al2O3-ZrO2/HZSM-5 catalysts. Catal Commun 5:367–370
93. Aguayo AT, Ereña J, Mier D et al (2007) Kinetic modeling of dimethyl ether synthesis in a
single step on a CuO-ZnO-Al2O3/γ-Al2O3 catalyst. Ind Eng Chem Res 46:5522–5530
130 F. Arena et al.

94. Aguayo AT, Ereña J, Sierra I et al (2005) Deactivation and regeneration of hybrid catalysts in
the single-step synthesis of dimethyl ether from syngas and CO2. Catal Today 106:265–270
95. Gao W, Wang H, Wang Y et al (2013) Dimethyl ether synthesis from CO2 hydrogenation on a
La-modified CuO-ZnO-Al2O3/HZSM-5 bifunctional catalysts. J Rare Earths 31:470–476
96. Bonura G, Cordaro M, Spadaro L et al (2013) Hybrid Cu-ZnO-ZrO2/H-ZSM5 system for the
direct synthesis of DME by CO2 hydrogenation. Appl Cat B 140–141:16–24
97. Wang S, Mao D, Guo X et al (2009) Dimethyl ether synthesis via CO2 hydrogenation over
CuO-TiO2-ZrO2/HZSM-5 bifunctional catalysts. Catal Commun 10:1367–1370
98. An X, Zuo Y-Z, Zhang Q et al (2008) Dimethyl ether synthesis from CO2 hydrogenation on a
CuO-ZnOAl2O3-ZrO2/HZSM-5 bifunctional catalyst. Ind Eng Chem Res 47:6547–6554
99. Aasberg-Pettersen K, Nielsen CS, Dybkjær L, Perregaard J (2009) Large scale methanol
synthesis from natural gas. www.topsoe.com. Accessed 24 Jun 2013
100. Lee S, Speight JG, Loyalka SK (2007) Handbook of alternative fuel technologies. Taylor &
Francis, Boca Raton
101. Dybkjær I (2006) Synthesis gas technology. www.topsoe.com. Accessed 21 Jun 2013
102. Halstead K (2008) Oryx GTL from conception to reality. www.fwc.com. Accessed
21 Jun 2013
103. Petroleum Technology Quarterly (2008) www.alfalaval.com. Accessed 23 Jun 2013
104. Sørensen EL, Perregaard J (2004) Condensing methanol synthesis and ATR – the technology
choice for large-scale methanol production. Stud Surf Sci Catal 147:7–12
105. Fuad MNM, Hussain MA, Zakaria A (2012) Optimization strategy for long-term catalyst
deactivation in a fixed-bed reactor for methanol synthesis process. Comput Chem Eng
44:104–126
106. Tijm PJA, Waller FJ, Brown DM (2001) Methanol technology developments for the new
millennium. Appl Catal A 221:275–282
107. Jahanmiri A, Eslamloueyan R (2002) Optimal temperature profile in methanol synthesis
reactor. Chem Eng Commun 189(6):713–741
108. Rahimpour MR, Alizadehhesari K (2009) Enhancement of carbon dioxide removal in a
hydrogen-permselective methanol synthesis reactor. Int J Hydrogen Energ 34:1349–1362
109. Van-Dal ES, Bouallou C (2012) CO2 abatement through a methanol production process.
Chem Eng Trans 29:463–468
Chapter 6
Syngas Production Using Carbon Dioxide
Reforming: Fundamentals and Perspectives

Julian R.H. Ross

6.1 Introduction

Much has been written over the last few decades about the potential use of CO2 as a
feedstock in the chemical industry, and particular attention has been given to its
possible use in the conversion of methane and other hydrocarbons using so-called
dry reforming to give syngas, a mixture of CO and hydrogen. The CO2 or dry
reforming of methane can be represented by the equation

CH4 þ CO2 ! 2CO þ 2H2 ð6:1Þ

In any practical application, in order to achieve acceptable conversions, this


reaction would have to be carried out at very high temperature (up to as much as
1,000  C) when the conversion would be close to chemical equilibrium. The
reaction is always accompanied by the reverse water–gas shift reaction, generally
fully at equilibrium:

CO2 þ H2 ⇄ CO þ H2 O ð6:2Þ

A consequence of the combination of these two reactions is that the CO:H2 ratio
obtained is generally rather lower than the value of 1:1 to be expected if only the
reaction of Eq. 6.1 occurred. The reaction system is further complicated by the fact
that the conditions used generally favour carbon deposition. Carbon deposition can
be thought of as occurring by one of three interdependent reactions:

2CO ⇄ C þ CO2 ð6:3Þ

J.R.H. Ross (*)


Charles Parsons Initiative, Department of Chemical and Environmental Sciences,
University of Limerick, Limerick, Ireland
e-mail: julian.ross@ul.ie

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 131
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_6,
© Springer-Verlag Berlin Heidelberg 2014
132 J.R.H. Ross

CH4 ⇄ C þ 2H2 ð6:4Þ

and

CO þ H2 ⇄ C þ H2 O ð6:5Þ

In most practical cases, whether or not carbon deposition is likely to occur is


determined by thermodynamic considerations; further, as Boudouard carbon
(Eq. 6.3) is generally expected at low temperatures, carbon deposition under dry
reforming conditions is most likely to occur by either methane decomposition
(Eq. 6.4) or CO reduction (Eq. 6.5). The problem of carbon deposition is a key
feature of the process and we will return to it many times below.
A search of Web of Science using the search terms “methane, CO2 and
reforming” shows that more than 1,900 papers with these three words in the titles
and/or abstracts had been published between 1987 and mid-2013. Table 6.1 shows
the approximate number of papers that include these search terms published up to
1994 and in each 5-year period since then, together with the approximate total
number of citations to these papers given in each period. It is clear that there is little
sign of the interest in the subject showing any decline; indeed, the yearly number of
publications reached an all-time high of more than 200 in 2012, this being signif-
icantly higher than the previous maximum of 160 in 2011. As will be discussed
further below, it is highly questionable if this level of activity is justified by the
potential applicability of the process. The only justification for such a level of
activity is probably that the work reported provides new insights into particular
types of catalysts, CO2 reforming being used as a model reaction.
Other combinations of search terms to those used in the search of the Web of
Science discussed above show up significant additional material.1 This review is
based largely on the material gleaned from the search related to Table 6.1. How-
ever, some references are also included that did not show up in this search, in most
cases because the keywords of the original articles were different; there is also no
doubt that some important papers will have been omitted for the same reason.
Interestingly, the first paper on CO2 reforming of methane listed in the search
outlined in Table 6.1 is a relatively under-cited one by Fish and Hawn on the use of
CO2 reforming in a thermochemical cycle [1], a topic to which we will return briefly
below. The second paper on the list, by Gadalla and Bowers [2], is concerned with
the use of a range of commercial steam reforming catalysts, showing that the most
promising results were obtained with alumina supports containing magnesium or
calcium oxide. (A further paper by Gadalla and Somer [3], published in 1989, gives
more details on this subject.) As will be discussed in a later section, Ni catalysts
based on Mg-containing supports still appear to be among the most effective
catalysts for the process.

1
No attempt has been made to cross-check the contents of various searches that have been carried out
so that the total number of papers is likely to be well in excess of 2000. More detailed Web of Science
and Scopus searches have also been carried out in order to identify some of the key references on each
of these subjects, and the text that follows below concentrates predominantly on these papers.
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 133

Table 6.1 The numbers of papers on the topic of dry reforming of methane published during
various time periods since 1987 together with the numbers of citations listed in those time periods
Years Up to 1995 1996–2000 2001–2005 2005–2010 2011 and later
Publications 60 280 490 670 >400
Citations <200 2,400 7,200 1,890 1,380
Data obtained from a Web of Science (Thomson Reuters) search (May 2013) using the search
terms methane, CO2 and reforming

Many of the papers that have been published over the last 20 years or so justify
the work carried out by claiming that a process involving the consumption of both
methane and CO2 would help to provide a solution to the emission of these two
greenhouse gases. It needs to be emphasised from the outset that such claims are
largely unjustified since the quantity of CO2 that might be used in syngas produc-
tion by CO2 reforming would be negligible compared with the very large emissions
of CO2 currently causing concern, even if it was possible to collect the CO2 in a
sufficiently pure form for the CO2 reforming reaction. Further, dry reforming
produces syngas with a composition usable only in limited applications; as shown
above (see Eq. 6.1), using methane as co-reactant, the syngas ratio, CO/H2, is
approximately 1.0. Much more suitable ratios can be attained by the addition of
either steam or oxygen to the reactant stream, such addition also having the benefit
that the potential for carbon deposition on the catalysts used in the process is
significantly reduced.2 An alternative to the dry reforming process using CO2 in
the primary feed is to operate the steam reforming reaction:

CH4 þ H2 O ! CO þ 3H2 ð6:6Þ

and then to carry out the reverse water–gas shift reaction (reverse of reaction (6.2))
in order to achieve the desired CO/H2 ratio:

CO2 þ H2 ! CO þ H2 O ð6:7Þ

There are several advantages to be achieved using this approach, these including
the following: (i) the problem of C deposition can largely be avoided since the
catalysts and process conditions for the steam reforming reaction under carbon-free
conditions are well established commercially; and (ii) the CO2 required for the
process can be extracted from the effluent of the reformer furnace (burning natural
gas) used to heat the tubular steam reforming reactor system. Hence, the only
additional processing step required is the separation unit to remove CO2 from the
combustion gases, a step also required if dry reforming is to be carried out. Further,
although the cost of a steam reforming plant is very high and a major contributor to
the all-over cost of any process requiring a syngas feed, it is unlikely that there will
be much reduction in the total operation cost if dry reforming is used.

2
A relatively small proportion of the papers in the recent literature have recognised these further
limitations, and some of these are listed in the tables below.
134 J.R.H. Ross

Accepting, for the reasons given above, that dry reforming is unlikely to solve the
greenhouse gas problem, other justifications have to be found for the very extensive
work on the subject that has been carried out over recent years. As will be discussed in
more detail in the following sections, much of the most recent work has been devoted
to trying to find catalysts that are resistant to carbon deposition while having suitable
activities and stabilities under reaction conditions.3 Although much of the work
published recently undoubtedly adds incrementally to the literature of the subject,
there appear to have been no major breakthroughs: most, if not all, of the catalysts that
have been examined under suitable conditions unfortunately show gradual deactiva-
tion due to carbon deposition. Further, many of the catalyst systems studied bear
remarkable similarity to ones that have been studied previously for other closely
related reactions, for example, the steam reforming of methane. In many cases, such
similarities do not appear to have been recognised by the authors of these papers.
This review traces briefly the development of processes used for the production of
syngas from methane (and higher hydrocarbons) and discusses the interest in the use
of CO2 as a co-reactant, paying particular attention to the conversion of methane. It
then considers the thermodynamics of such conversion processes, highlighting the
problem of carbon deposition and the need for operation at high temperature.
Recognising the similarity between dry reforming and steam reforming and the fact
that the latter is a well-established industrial process, mention is made of the catalysts
used commercially for the steam reforming of methane and of higher hydrocarbons,
with a short digression on the catalysts used for methanation. The review then
considers various groups of papers that have been devoted to studies of different
catalyst types, particular attention being given to the use of noble metals (particularly
Pt) and other transition metals such as Ni, Co and Mo, on a variety of different
supports. The review concentrates on the literature on methane reforming, but it
should be recognised that many of the catalysts developed will be equally applicable
to the reforming of higher hydrocarbons, the only complication being that methane is
also a potential product when the methanation reaction is thermodynamically feasible
(at lower temperatures, see Sect. 6.2.2).

6.2 Historical Background

6.2.1 Syngas Production from Hydrocarbons

Before embarking on a more detailed discussion of the dry reforming reaction and
the catalysts that have been used for that reaction, it is sensible first to give a brief

3
Unfortunately, a significant proportion of the papers in the recent literature report experiments
carried out under unrealistic conditions, for example, at low temperatures or with
non-stoichiometric reactant compositions; in any process that will be developed, the reaction
will be carried out at high temperatures and pressures under conditions close to equilibrium, and
any novel catalyst formulations must be tested under similar conditions.
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 135

outline of the use and importance of the steam reforming of methane as a source of
syngas. The subject of the steam reforming of methane (Eq. 6.6) has been reviewed
by various different authors, from the point of view of both the processes involved
[4] and of the catalysts used [5]. Although, as pointed out by Rostrup-Nielsen [4],
some papers and patents on the steam reforming of hydrocarbons had been
published as early as 1868, the first industrial steam reformer was commissioned
in Baton Rouge in 1930 by the Standard Oil of New Jersey, and this was followed
by another reformer commissioned by ICI in Billingham in 1936; in both cases, the
main product required was hydrogen, and so the CO formed was removed using the
water–gas shift reaction. While steam reforming in the United States was usually
carried out with methane as the feed hydrocarbon, the raw material of choice in
Europe until the 1960s was naphtha; with the advent of European natural gas
supplies in the second half of that decade, methane also became the feedstock of
choice.4 It is interesting to note that although there has been some limited com-
mercial interest in the use of CO2 as a feed instead of water, the only plant using a
methane/CO2 feedstock has been a pilot plant operated by Haldor Topsoe in Texas
in which a sulphur-modified steam reforming catalyst is used [6]. It is also worth
noting that in the so-called Midrex process for the direct reduction of iron ore, a
mixed feed of methane with CO2 and steam is used in a reformer; the CO2 and
steam for the reforming step are produced during the reduction of the iron ore, and
the syngas produced in the reformer is fed directly to the reduction reactor [6, 7].
Although other active materials, notably the noble metals, have been examined
for use as the active components in steam reforming catalysts, the only ones that
have been used commercially are those including Ni as the active component. This
is not only a matter of cost: Ni is the only non-noble transition metal that is
maintained in its metallic state under steam reforming conditions since the
equilibrium

NiO þ H2 ⇄ Ni þ H2 O ð6:8Þ

lies to the right-hand side as long as the ratio H2/H2O is greater than about 0.01; in
other words, the nickel will be in its reduced (and active) form as long as there is
approximately 1 % hydrogen in the reactant gas [5]. (With the exception of the
noble metals, much higher proportions are needed for most of the other transition
metals.) A variety of different Ni-containing formulations have been
commercialised. However, among the most commonly used materials are the ICI
(now Johnson Matthey) 46–1 catalyst (predominant components in the reduced
form are Ni, alumina and K2O with CaO, SiO2 and MgO added as a cement binder)
and the Haldor Topsoe catalyst (containing Ni, MgO and alumina). Other catalyst
manufacturers supply similar materials. The potassium of the ICI formulation is
added to help minimise C deposition, and the MgO of the Haldor Topsoe material

4
Syngas could also be formed by coal gasification or by modifications of the gasification process
using steam or CO2 in the feed.
136 J.R.H. Ross

serves a similar function. As mentioned above, the catalyst used in the Haldor
Topsoe pilot plant for dry reforming (see above) is partially sulphided [4, 6, 8]. This
was based on work on the subject of S passivation for application in the steam
reforming process [9].

6.2.2 Methanation

From a historical point of view but also in relation to work on the formulation of
catalysts, it is interesting to note that there occurred in the late 1970s and early
1980s a period of very significant academic and industrial research activity on the
topics of methanation and methanation catalysts. This interest arose because natural
gas supplies in the United States could not keep up with demand. As a result, it was
desirable to find means of converting syngas (produced from coal) into so-called
synthetic natural gas (SNG):

CO þ 3H2 ! CH4 þ H2 O ð6:9Þ

This reaction is the reverse of the steam reforming reaction (Eq. 6.6) and is highly
exothermic. Hence, much of the work from that period concentrated on finding
catalysts with high stability at higher temperatures that also resisted carbon deposi-
tion. Many of the catalysts examined and reported in many papers of that period had
great similarity to those that have been examined more recently for dry reforming.
This author reviewed the literature on methanation and steam reforming in 1985 [10].

6.2.3 Methane Coupling

Following on from the great interest from the catalysis community in the topic of
methanation, the subject of the oxidative coupling of methane to give C2H6 and
C2H4 became predominant during the second half of the 1980s as a result of a paper
by Keller and Bhasin published in 1982 [11]. These authors showed that it was
possible, using a cyclic process in which methane and oxygen were fed in turn to a
catalyst bed, to obtain reasonable yields of C2 hydrocarbons, the catalysts being
operated at temperatures around 800  C. Keller and Bhasin used a range of
α-alumina-supported materials (including the oxides of Cd, Sn, Sb, Tl, Bi and Pb)
as catalyst and postulated that an oxidation–reduction mechanism was involved.
Hinsen and Baerns were the first to carry out the reaction with methane and oxygen
fed simultaneously, and they obtained good yields using a lead/γ-alumina catalyst
[12]. A paper by Ito, Wang, Lin and Lunsford published in 1985 introduced the
use of Li/MgO materials [13]. Many more similar catalysts were introduced over
the next decade, and a number of very extensive reviews have been written on the
subject [14–17].
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 137

It was soon recognised that the reaction occurs by a combination of heterogeneous


reactions and gas-phase radical reactions and that the products included not only
ethane and ethylene but also a mixture of higher hydrocarbons, water, CO, CO2 and
hydrogen. The participation of gas-phase reactions as well as limitations introduced
by the explosive limits of methane/oxygen mixtures leads to a limit in the potential
yields of C2 hydrocarbons produced in the reaction, the best values being just above
20 %, too low to enable the reaction to have commercial application with current
natural gas prices. Although papers on the subject were published in a variety of
different journals, a very good overview of the work that was carried out is to be
found by examining the proceedings of the earlier “Natural Gas Conversion” meet-
ings [18–21], from the first of the series held in 1987 in Auckland, NZ, to the fourth
meeting held in 1998 in Messina, Italy; by 2001, when the 6th meeting of the series
was held in Alaska, the topic of methane coupling had disappeared. What had been
gained from all this work was a much greater understanding in the academic
community of the operation of catalytic reactors at elevated temperatures and the
problems of catalyst stability. (It is of interest to note that renewed interest in methane
coupling has recently been evident, this being spurred by the current low prices of
natural gas.)

6.2.4 Dry Reforming and Partial Oxidation of Methane

When at the beginning of the 1990s an interest developed in the production of


syngas from natural gas using either partial oxidation or “dry reforming”, it was not
surprising that the research community that had been involved in the work on
methane coupling transferred its interest to studying one or other of these two
reactions. This transfer is clearly seen in the proceedings of the “Natural Gas
Conversion” meetings referred to above. The majority of the papers under the
title “Methane Conversion” in the second meeting held in Oslo in 1990 [18] were
related to oxidative coupling, there being none on partial oxidation or CO2
reforming. This bias was continued at the third meeting held in Sydney in 1993
[19] with more than 20 papers on methane coupling and only four on dry reforming;
two of the latter were reviews of the subject by J. R Rostrup-Nielsen and by J. H.
Edwards and A. M. Maitra (discussed further in the following section). However,
by the time of the fourth meeting held in the Kruger National Park in South Africa
in 1995 [20], there were 16 papers on oxidative coupling, 13 papers on partial
oxidation and 8 papers on CO2 reforming. By the fifth meeting held in Taormina,
Italy, in 1998 [21], the latter two topics had become the predominant subjects of
interest related to methane conversion. The relevant papers from these meetings
will be outlined in Sect. 6.4.1.
As discussed above (see Table 6.1), there has since 1996 been an explosive
development of interest in CO2 reforming and partial oxidation, so much so that it is
virtually impossible to write a comprehensive review covering every aspect of the
work that has been carried out on either topic. The following sections will therefore
138 J.R.H. Ross

concentrate on some of the more important types of catalyst that have been
investigated, particular attention being paid to papers describing those types of
catalyst that have reasonable stabilities under “dry reforming” conditions and to
papers giving new insights into the mechanism of the reaction or advancing other
approaches to the conversion process. No great detail on any of the papers is given;
instead, summaries are given in tabular form of their main conclusions. Further, no
attempt is made to discuss any of the parallel literature on partial oxidation,
exceptions being when it has been shown that the addition of oxygen to the “dry
reforming” mixture has beneficial effects or when the results on the partial oxida-
tion reaction on a particular catalyst have some relevance to the dry reforming
reaction on identical or similar catalysts. In selecting the literature for inclusion, use
has been made of citation indexing, the searches being based on the use of Web of
Science and/or Scopus. As a starting point, some of the papers appearing in the
proceedings of the earlier Natural Gas Conversion meetings referred to above have
been used to identify some of the major themes of research. Scopus and Web of
Science searches have then been used to identify some of the (in most cases) more
recent key papers related to each of these themes, and some of the main conclusions
of these are tabulated. Finally, the main objectives and conclusions of some very
recently published papers on the subject that have not yet achieved a significant
citation record are summarised.
Before embarking on our summary of the literature, a brief outline of the
thermodynamic limitations associated with the dry reforming reaction and associ-
ated side reactions is given since these limitations have some importance in
discussing the papers from the literature.

6.3 Thermodynamics of the CH4 + CO2 Reaction

The dry reforming of methane (Eq. 6.1) is an endothermic process, and so the
maximum conversion calculated thermodynamically increases with increasing tem-
perature. This is illustrated schematically in Fig. 6.1 that shows the thermodynamic
conversions for various feed compositions as a function of temperature of reaction
[22]. It can be seen that a temperature in excess of about 850  C is necessary to obtain
adequate conversions using a stoichiometric CH4/CO2 feed (1:1).
It is to be expected that the water–gas shift reaction (Eq. 6.2) will be in
equilibrium under CO2 reforming conditions and so the CO/H2 ratio of the product
gas can be calculated easily from a knowledge of the feed compositions and the
measured methane conversions. It should be noted that the so-called selectivity of
the reaction is very frequently given in publications reporting the behaviour of
novel catalysts; unless it is shown very clearly that the water–gas reaction is not in
equilibrium for that catalyst system, such values have no real significance as they
are thermodynamically rather than kinetically controlled.
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 139

Fig. 6.1 Thermodynamically calculated conversions of methane as a function of temperature for a


series of different feed ratios. Calculations carried out using HSC chemistry (version 1.10,
Outokumpu Research, Finland) (Reprinted from Ref. [22]. Copyright 2005, with permission
from Elsevier)

Figure 6.2 shows the yield of carbon to be expected if carbon deposition is


possible for two possible cases: whether or not the water–gas shift reaction is
considered to be in equilibrium [22]. This diagram was constructed assuming that
there is a closed system and that the amount of carbon formed relates to the amount
fed as methane and CO2; the results should therefore be seen only as a guide as to
when carbon would form in a continuous flow reactor. It is clear that carbon
formation is possible over the whole range of temperatures in both cases considered
(with or without the water–gas shift reaction being equilibrated), and that the
amount of carbon formed will be lower at high temperatures. When the water–gas
shift reaction does not take place, the amount formed at the highest temperatures is
much lower; however, in most cases reported in the literature, the water–gas shift
reaction is at equilibrium at the higher temperatures and so this situation does not
apply. Rostrup-Nielsen uses a different approach, favouring the use of so-called
carbon limit diagram, developed first for the steam reforming reaction [6, 23]. He
shows that in the absence of the addition of steam, the CO2 reforming reaction
operates under conditions when carbon can form regardless of the CH4/CO2 ratio.
He concludes that the noble metals, most of which do not form carbides (necessary
intermediates in carbon formation) have a greater potential for use as CO2
reforming than do catalyst formulations using nickel. The topic of carbon deposi-
tion will be handled further after a discussion of the types of catalyst studied.
140 J.R.H. Ross

Fig. 6.2 Thermodynamically calculated proportions of carbon formed under CO2 reforming
conditions with CH4/H2O ¼ 1.0; (■), with the reverse water–gas shift reaction in equilibrium;
(□), without the reverse water–gas shift reaction (Reprinted from Ref. [22]. Copyright 2005, with
permission from Elsevier)

6.4 Catalysts for the Dry Reforming of Methane

6.4.1 Papers on CO2 Reforming Published


in the Proceedings of the Natural Gas Conversion
Symposia up to 1998

Tables 6.2, 6.3, and 6.4 list the relevant papers presented at the Natural Gas
Conversion Symposia between 1993 and 1998: the third (held in 1993, published
in 1994 [19]), the fourth (held in 1995, published in 1997 [20]) and the fifth (held
and published in 1998 [21]). The following paragraphs first summarise two review
articles from the third meeting and then outline some of the important aspects of the
other papers, this being in preparation for a more general discussion of the literature
on the subject since about 1997.
In his review delivered at the third symposium [6], Rostrup-Nielsen of Haldor
Topsoe A/S (Lyngby, Denmark) gave a very clear exposition of the then-available
information on the CO2 reforming of methane, paying particular attention to earlier
work on the subject. He noted that the reaction was examined by Fischer and
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 141

Table 6.2 Papers from NGC 3


Temperature
Reference/year Metal Support range/ C Comments
Rostrup-Nielsen [6] Noble metals Alumina No specific Review article, discussing
S-passivated MgO range process options and
nickel given carbon deposition; also
summarises some work
from Haldor Topsoe on
noble metals as CO2
reforming catalysts
Edwards [24] Range Range No specific Review article, summarising
range types of catalysts studied
given up to about 1993; some
discussion of possible
processes
Seshan [25] Pt, Ni ZrO2, Al2O3 390–560 Pt/ZrO2 much more stable
that Pt/Al2O3; 5%Ni/
ZrO2 more stable than
10%Ni/ZrO2
Uchijima [26] Rh Al2O3, SiO2, 377–727 PO gives first combustion
TiO2 and then steam and CO2
reforming. Rh/Al2O3 is
35 more active than
Rh/SiO2 for CO2
reforming. Addition of
Al2O3 or TiO2 to Rh/SiO2
gives significant
promotion

Tropsch as early as 1928 [49]; these scientists studied a series of base metal
catalysts and found that nickel and cobalt gave the best results, the product gas
compositions being close to thermodynamic equilibrium.
Rostrup-Nielsen also summarises the work of Bodrov and Apel’baum who showed
in 1967 [50] that the kinetic expression for the CO2 reforming reaction over a nickel
film was similar to what they had found previously for the steam reforming reaction
[51], this indicating that the mechanism was very similar for both reactions.
Rostrup-Nielsen discussed industrial processes that involve CO2 reforming,
these including the SPARG process making use of a sulphur-passivated Ni catalyst.
The original research relating to that process is detailed in a research paper by
Dibbern et al. referred to above [8]. The technology used, adding traces of H2S to
the reactor feed, is similar to that previously introduced by Haldor Topsoe to enable
steam reforming to be carried out at low steam/methane ratios [9]. Rostrup-Nielsen
also presented in his review some data on the CO2 reforming of methane using the
noble metals and Ni supported on MgO. He concluded that the noble metals are
much less susceptible to carbon poisoning due to the fact that they do not dissolve C
in their bulk. He then described briefly some work in which Rh and Ru catalysts, the
142 J.R.H. Ross

Table 6.3 Papers from NGC4


Reference/ Temperature
year Metal Support range/ C Comments
Cant [27] Ni, Rh Al2O3 700 Deuterium kinetic isotope effect
shows differences between metals.
C easily removed from Rh
Slagtern Ni Al2O3 700, 800 Rare earth mixture used as promoter;
[28] similar results to using La2O3
Yu [29] Ni Al2O3 700, 750 CH4 decomposition, followed by
removal of C by CO2
Zhang [30] Ni La2O3 550–750 L2O3 activates CO2 and encourages
removal of C from Ni
Erdöhelyi Ir Al2O3, TiO2, 427–500, C deposition limited on all samples.
[31] SiO2, MgO 850 TiO2 most effective support
Van Keulen Pt ZrO2 550–800 Optimisation of Pt/ZrO2 with life tests
[32]
Kikuchi [33] Ni, noble Al2O3 500 Tested in membrane reactor system.
metals Noble metals much less susceptible
to coke formation
Ponelis [34] Ru, Rh SiO2 400–900 Membrane gives significantly
improved conversions at higher
contact times

most promising of the metals studied, were used for carbon-free operation in a pilot
plant under conditions in which carbon is favoured thermodynamically. He con-
cluded, however, that because of their scarcity neither of these metals is likely to be
used in other than niche applications. In conclusion, Rostrup-Nielsen discussed the
limitations of possible processes for the use of CO-rich gases from CO2 reforming.
He pointed out that the amounts of CO2 used in such processes will be virtually
insignificant in comparison with the total worldwide emissions of CO2. For exam-
ple, the CO2 that would be required for the production of 5 million tons of acetic
acid per year (the current global rate of production) by the reaction:

CH4 þ CO2 ! CH3 COOH ð6:10Þ

would correspond to the emission from only one 500 MW coal-based electricity
power station. Yearly worldwide production of methanol is four times higher, and
so supplying syngas via CO2 reforming of methane would require the output from
four such power plants. More importantly, the current technology for CO2 extrac-
tion makes the use of CO2 from flue gases uneconomical and so other sources of
CO2 would have to be considered. The article by Rostrup-Nielsen contains many
other details (and warnings) related to the potential of a process for CO2 reforming.
Anyone working in the area should therefore read it with care.
A second review article was delivered at the same symposium; this was reported
by Edwards and Maitra of CSIRO (North Ryde, Australia) [24]. Having also
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 143

Table 6.4 Papers from NGC V, SSSC 119 (1998) (Italy)


First author/ Temperature
reference Metal Support range/ C Comments
Hallische [35] Ni with Co, Ce, α-Al2O3 550, 730 Co, Ce, Cu promote
Cu or Fe as activities
promoter H2/CO>1
Buarab [36] Ni Various supports 600 α-Al2O3 most effective
support; K poisons
reaction
Nichio [37] Ni α-Al2O3 650–800 Best results for preparation
with Ni acetylacetonate
Gronchi [38] Ni SiO2, La2O3 600 Pulse reactor study
CO2 reacts with La2O3
Provendier [39] Ni LaNiFe perovskite 400–800 Perovskite forms Ni–Fe
alloy in use. Low coke
formation
Kroll [40] Ni SiO2 700 Detailed kinetic study
Kim [41] Ni Al2O3 600–800 Aerogel catalyst gives low
carbon deactivation
K addition helpful
York [42] Ni + W or Mo Al2O3 700 W improves Ni catalyst
oxides
Mo2C, WC No support 850 Carbides vs. stable
Suzuki [43] Ni Ca0.8Sr0.2TiO3 850 In situ preparation. Stable
BaTiO3 against C formation
Stagg [44] Pt ZrO2 with 800 Promoters increase
CeO2, La2O3 stability and activity
O’Connor [45] Pt ZrO2 600 DRIFTS study; oxidation
–reduction mechanism
probable
Quincoces [46] Ni CaO/SiO2 650 CaO increases activity
Nam [47] Ni Ni–La perovskites 600–900 Addition of Ca reduces
with Sr coke formation
promotion
Tomishige [48] Ni MgO from solid 500–850 Addition of small quanti-
solution ties of H2 stops any
deactivation

discussed the various reactions involved and their thermodynamics, these authors
gave a very comprehensive review of work on CO2 reforming published up until
1993. They showed that most of the noble metals and also nickel had been studied
and that oxides such as alumina and magnesia had been used as supports for the
active metals. They also pointed out that carbon deposition rates seem to depend on
the support used and possibly on catalyst morphology, two factors that emerge
repeatedly in more recent publications. They mentioned specifically the paper by
Gadalla and Somer referred to above [3] in which the use of a Ni/MgO catalyst for
up to 125 h was reported; at the end of the experiment, these authors observed only
a minute trace of carbon deposition.
144 J.R.H. Ross

Edwards and Maitra also discussed the results of two papers where the activities
of a series of metals are reported. In the first, by Takayasu et al. [52], Ni, Ru, Rh, Pt
and Pd catalysts supported on “ultrafine single crystal MgO” were compared, the
activities being in that order. In the second, by Ashcroft et al. [53], catalysts
comprising of Ni, Ir, Rh, Pd and Ru supported on alumina were examined, the
order of activity being in the order given. While Takayasu et al. had found that Ru
was one of the most active metals, Ashcroft et al. reported that it is one of the least
active. (Rostrup-Nielsen has also reported it as being very active [6].) However,
both groups found that Ni and Rh are among the most active metals. Listing other
papers on the topic, Edwards and Maitra also pointed out that Solymosi et al. [54]
and Masai et al. [55] had previously reported that Rh supported on alumina is one of
the most active catalysts for the reaction.
Edwards and Maitra closed their review [24] with a discussion of the potential
application of thermochemical heat pipe applications for the recovery, storage and
transmission of solar energy, showing that the CO2 reforming reaction can be
coupled with the methanation reaction to provide a method of energy transmission
(a topic also mentioned briefly by Rostrup-Nielsen in his more extensive review
[6]). We return briefly to this topic later (see Table 6.13).
The research papers on CO2 reforming of methane from the meetings of the
Natural Gas Conversion series listed in Tables 6.2, 6.3, and 6.4 cover the use of
most of the metal/support combinations that have since been examined in more
detail. The vast majority of these papers reported work involving nickel, this being
supported on a variety of different oxides, in particular, alumina [6, 25, 27–29, 33,
35, 37, 41, 42], silica [31, 38, 40, 46], magnesia [6], lanthana [30, 38] and zirconia
[25]. Some attention was also given to the use of catalysts derived from nickel-
containing compounds or from solid solutions of NiO in MgO [48]. Most of the
noble metals are also featured, particular attention being focused on Pt [2, 25, 32],
Rh [26, 27, 34], Ru [34] and Ir [31], again on a variety of supports. It is interesting
to note that there was a wide range in the temperatures of operation, from 400 to
900  C. As was pointed out above, operation at high temperature is necessary to
give acceptably high conversions and so the results at lower temperatures, although
giving information on relative activities, have less practical relevance. It should
also be noted that the paper by York et al. [42] reports the use of carbides, while the
papers by Kikuchi and Chen [33] and Ponelis and Van Zyl [34] both describe the
use of membrane systems.

6.4.2 Highly Cited Publications from the General Literature

In the following sections, we give a listing of the more significant papers from the
recent literature under headings suggested by the brief outline given in the last
paragraph. Owing to the very large literature on the subject of CO2 reforming of
methane, the papers selected for mention under each heading are chosen in relation
to their citation records, recognising that those papers most frequently referred by
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 145

Table 6.5 Some of the most cited reviews on CO2 reforming listed in the Web of Science
First author/ Citations (WoS
reference Comments May 2013)
Wang [58] Identifies areas for future research: effect of metal-support 226
interaction on activity; effects of promoters; reaction kinetics
and mechanism; pilot reactor performance and scaleup.
Ni on some supports gives carbon-free operation
Hu [59] Reviews catalysts studied, with emphasis on inhibition of C 212
deposition (also reviews partial oxidation)
Ross [60] Review of the development of Pt/ZrO2 catalysts 134
Ross [61] General review of CO2 reforming 59
Bradford [62] Reviews the use of Ni and Rh catalysts, also single crystals; 552
also partial oxidation
Verykios [63] Review concentrating on the development of Ni/La2O3 catalysts.
La2O2CO3 species are critical to good operation
Armor [64] Review largely on steam reforming but also including CO2 445
reforming, particularly limitations due to C deposition and
also costs of supplying pure CO2
Ma [65] Review of CO2 conversion including CH4 + CO2 over Ni 45
supported on a wide variety of supports
Kodama [66] Review on production of fuels using sunlight including CO2 118
reforming of methane

others are likely to be the most important in that field. In the tables, the number of
citations to each of the publications given by the Web of Science in May 2013 is
shown for comparison purposes; clearly this is only a snapshot of the situation, and
it should also be recognised that younger publications will have correspondingly
lower numbers of citations than older ones.5 A reader with a particular interest in
one or another type of catalyst would gain an up-to-date picture of the literature on
that topic by doing his/her own citation study based on the appropriate key papers
listed in the tables.
Table 6.5 lists some of the most cited reviews on the subject of CO2 reforming.
Some of these are very general, covering a whole range of topics from basic aspects
to pilot plant results [58], while others are much more focused on one type of
catalyst [60, 63]. The review by Armor [64] is of particular interest as it not only
covers both steam reforming and CO2 reforming but also deals with matters such as
problems associated with purification of CO2. Kodama [66] reviews the subject of
fuel production using sunlight.
Nickel, although very active for the conversion of methane by either steam reforming
or CO2 reforming, suffers from the problem that it rapidly becomes deactivated by the
formation of carbon, this carbon generally being in the form of nanotubes with a nickel
crystallite at the tip. Much of the effort has therefore been in trying to find suitable
additives for the catalysts that help to minimise this form of carbon deposition, often by
modifying the Ni particle size or the support surface in the region of the nickel

5
We have paid some attention to this time dependence in our choice of which papers to include.
146 J.R.H. Ross

crystallites. It is now reasonably well established that small nickel crystallites are much
less susceptible to C growth, but anchoring the active metal to the support is also known
to hinder filament growth. Table 6.6 lists some of the papers reporting the use of
catalysts consisting of Ni supported on alumina or modified aluminas. A number of
additives are seen to improve the behaviour of the catalysts in relation to the problem of
carbon deposition, particularly the alkali [67, 75] and alkaline earth metals [67, 74],
lanthana [71] and ceria [68, 73]; ceria is particularly effective.
The effect of reducing the particle size is seen particularly with catalysts based
on Ni on MgO; see Table 6.7. This is because the precursor to the active catalyst
often involves the formation of a solid solution of NiO in MgO; see, for example,
references [77, 80, 83]. Catalysts based on supports such as MgAl2O4 are also very
effective [81–83].
Another support that has received some attention is zirconia, sometimes alone
but also often promoted by other species such as ceria [84, 85] or magnesia [83]; see
Table 6.8. The effect of Ni particle size on stability has been shown for a zirconia
support by Lercher et al. [86]. Other supports studied have included titania [88],
lanthana [89], silica [90] and a La–Sr–Ni–Co perovskite [91]; see Table 6.9.
Noble metals have the advantage over Ni in that they do not dissolve carbon, and
so the problem of the growth of carbon filaments does not occur with them. As a
result, there has been a significant amount of work using catalysts containing noble
metals on a variety of supports. Table 6.10 lists some highly cited papers describing
the use of supported Pt catalysts on various different supports. Work from this
laboratory [60, 95] and also from Bitter et al. [96] has been concerned with the use
of Pt on zirconia, this usually being promoted by a small quantity of alumina. Other
promoters such as La2O3 or CeO2 are also possible [93].
The other noble metals have also received significant attention, particularly Rh
[97–102], Ir [99, 101] and Pd [101, 103]; see Table 6.11. As discussed by Rostrup-
Nielsen, a series of successful pilot plant tests have been carried out using a catalyst
based on Rh. However, the cost of Rh is prohibitively high so that this work is
unlikely to be applied.
Because of the high cost of the noble metals and the resultant problems of
operating large scale using them, a significant amount of attention has been paid
to the possibility of using MoC2 or WC as the active component of a catalyst, it
being well established that these carbides exhibit properties similar to those of the
noble metals. Claridge et al. ([105], see Table 6.12) showed that both of the carbides
were active for the reaction, but they reported that there was some evidence of
deactivation caused by oxidation of the carbides. Work from this laboratory [106]
showed that this was probably due to the fact that the entrance to the CO2 reforming
bed operates under slightly oxidising conditions, this causing the carbide to oxidise,
and that the exit part of the bed was stable. The activity of a Mo2C material
supported on ZrO2 could be improved significantly by promotion with Bi species.
Finally, Table 6.13 lists some references describing work in various laboratories
on energy transport systems involving the CO2 reforming reaction as the endother-
mic part of the system. It is interesting to note that despite the comments above, the
metal of choice for this work appears to be Rh, probably because of the reliable
performance far outweighs the disadvantage of cost.
Table 6.6 Some of the most cited papers dealing with Ni supported on Al2O3 or modified Al2O3
First author/ Temperature Citations
reference Metal Support range/ C Comments (WoS May 2013)
Horiuchi [67] Ni Al2O3 Studied effect of adding Na, K, Mg and Ca. All decrease 185
the tendency to deposit carbon
Wang [68] Ni CeO2–Al2O3 CeO2 addition (1–5 wt% optimum) gives better activity, 182
stability and C suppression
Wang [69] Ni Al2O3, SiO2, MgO Strong metal-support interaction improves stability and 129
coking resistance
Kim [70] Ni Al2O3 aerogel 700 Aerogel supported metal more active and stable than con- 154
ventional Ni/Al2O3 catalyst. C deposition depends on
Ni particle size, being significant above 7 nm
Martinez [71] Ni Al2O3 Studies influence of La2O3 addition. La decreases rate of 79
coking
Juan-Juan [72] Ni, Ni/K Al2O3 TP reaction to 900, Coking rate with K much reduced 59
isothermal at 700
Laosiripojana [73] Ni Al2O3–CeO2 800–900 8 % CeO2 gives greatly improved behaviour, with greatly 54
decreased C deposition
Goula [74] Ni CaO–Al2O3 750 Ca content influences amounts of C on catalyst after use 90
and also Ni morphology. Origin of C probed by tracer
study
Osaki [75] Ni K–Al2O3 K controls Ni ensemble size and thus affects C deposition 93
rates
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives
147
148

Table 6.7 Some papers dealing with Ni supported on MgO


First author/ Temperature Citations (WoS
reference Metal Support range/ C Comments May 2013)
Rostrup-Nielsen Ni MgO Discussion of all aspects of the reaction over Ni catalysts 674
[76]
Tomishige [77] Ni MgO, Al2O3 Ni0.03MgO0.970 solid solution much more resistant to C 239
deposition than Ni/MgO or Ni/Al2O3
Wei [78] Ni MgO Careful kinetic analysis (with isotopic measurements) 225
shows C–H bond activation to be sole relevant step
Ruckenstein [79] Ni MgO and other alkaline Ni/MgO more stable than Ni on the other supports; MgO 194
earth oxides inhibits Boudouard reaction
Ruckenstein [80] Ni MgO 790 Combines CH4/O2 with CH4/CO2 to decrease chance of 59
SiO2 explosion. Ni/MgO solid solution catalyst most effective
Al2O3
Djaidja [81] Ni MgO/Al2O3 800 Ni/MgO sample prepared by co-precipitation less stable 58
MgO than one prepared by impregnation
Guczi [82] Ni MgAl2O4 Study of C species formed. Addition of Au shown to 18
improve activity and stop formation of C nanotubes
Wang [83] Ni, Co, Ni–Co MgAlOx 750 Ni–Co catalyst is very stable; no deactivation after 2,000 h
operation. XANES and EXAFS study of catalyst
structure
J.R.H. Ross
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 149

Table 6.8 Some papers dealing with Ni supported on zirconia


Citations
First (WoS
author/ Temperature May
reference Metal Support range/ C Comments 2013)
Roh [84] Ni Ce/ZrO2 750 Work on steam reforming of 152
methane using catalysts previ-
ously studied for CO2 reforming
Montoya Ni ZrO2–CeO2 500–800 CeO2 stabilises tetragonal ZrO2. 133
[85] Graphitic C does not cause
deactivation. CeO2 enhances
reverse water–gas shift reaction
Lercher Ni, Pt ZrO2 and other 600, Ni/ZrO2 very stable if low Ni 94
[86] supports 720–780 loadings used. (Ni particle size
~2 nm)
Nagaraja Ni K/MgO–ZrO2 550–750 Addition of 0.5wt%K increases 3
[87] activity and stability of the
Ni/MgO/ZrO2 catalyst

Table 6.9 Some papers dealing with nickel on other supports


Citations
(WoS
First author/ Temperature May
reference Metal Support range/ C Comments 2013)
Bradford Ni TiO2 (also 400–550 Ni/TiO2 more active than 231, 152
[88] SiO2, MgO) Ni/C, Ni/SiO2 or Ni/MgO.
Ni/TiO2 suppresses C
deposition. Reverse
water–gas shift reaction in
equilibrium
Zhang [89] Ni La2O3, Al2O3, 750 Sites at Ni–La2O3 interface 157
CaO/ Al2O3, give active and stable per-
CaO formance compared with
other supports all of which
gave continuous
deactivation with time
Swaan [90] Ni SiO2 700 Operating catalyst is Ni 181
carbide. Deactivation by
catalyst sintering and C
deposition
Valderrama Ni–Co La–Sr–O 600–800 La–Sr–Ni–Co–O perovskite 23
[91] solid solutions give rise
to nano-sized crystallites
stable under reforming
conditions
150

Table 6.10 Some papers dealing with Pt catalysts on various supports


Citations
First author/ Temperature (WoS May
reference Metal Support range/ C Comments 2013)
Ashcroft [92] Pt and other Al2O3 400–800 Shows that partial oxidation and CO2 reforming can be 437
Pt- group metals carried out simultaneously
Stagg-Williams [93] Pt ZrO2, ZrO2/La2O3, 800 La stabilises support. Ce improves oxygen uptake of 124
ZrO2/CeO2 support
Damyanova [94] Pt CeO2–Al2O3 450–600 CeO2 gives improved Pt dispersion and increased 118
activity. Optimum 1 % CeO2
Ross [60] Pt ZrO2, TiO2, Al2O3 550–800 Pt on ZrO2 much more stable with respect to C 130
deposition than other samples
O’Connor [95] Pt ZrO2, Al2O3 550–800 O2 addition to CO2 reduces the temperature for a given 99
conversion and also reduces carbon deposition
Bitter [96] Pt ZrO2, SiO2 600–800 ZrO2 as support activates CO2, while Pt black or SiO2 113
as a support does not
J.R.H. Ross
Table 6.11 Some papers dealing with catalysts containing Rh, Ru, Ir or Pd
First author/ Citations (WoS
reference Metal Support Temperature range/  C Comments May 2013)
Nakamura [97] Rh Al2O3, TiO2, SiO2 277–727 Rh/SiO2 greatly improved by mixing with 109
Al2O3, TiO2 or MgO
Richardson Rh, Pt/Re γ-Al2O3 washcoated on Pt–Re deactivates Foams exhibit high effectiveness factor. Rh 54
[98] α-Al2O3 below 700  C best overall catalyst
Mark [99] Rh, Ir Range of oxides 550–850 Specific rates constant regardless of support. 129
H2/CO ratio varied by addition of
H2O or O2
Erdohelyi [100] Rh Al2O3, TiO2, SiO2, MgO 150 upwards Methane adsorbs above 150  C. Reaction with 29
CO2 occurs above 400  C. No carbon
deposition observed
Erdohelyi [103] Pd Al2O3, TiO2, SiO2, MgO 400–500 Little carbon deposited with stoichiometric gas 127
mixture
Qin [101] Ru, Rh > Ir > MgO 600–900 Compares steam and CO2 reforming. Activity 121
Pt, Pd in order shown. Ru, Rh and Ir give little
deactivation. Mixed reforming slower
Maestri [102] Rh Al2O3 300–900 Microkinetic analysis of data of Donazzi 44
et al. [104]
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives
151
152 J.R.H. Ross

Table 6.12 Some papers dealing with carbides as catalysts


Citations
First author/ Temperature (WoS May
reference Carbide Support range/ C Comments 2013)
Claridge [105] Mo2C, Al2O3, ZrO2, 850–950 Interaction with support 157
unsupported SiO2, TiO2 during preparation
WC affects activity. Oxi-
dation of carbide
causes deactivation
Treacy [106] Mo2C Al2O3, ZrO2, 950 ZrO2 material most stable. 2
SiO2 Addition of Bi gives
higher conversions.
Catalyst oxidises
at entry to bed

Table 6.13 Some papers discussing the use of dry reforming as part of an energy transportation
system
Citations
(WoS
First author/ Temperature May
reference Metal Support range/ C Comments 2013)
Fish [1] – – – Describes balancing of the 27
reactions for a system
including dry
reforming (Sandia
Natl. Labs.)
Levy [56] Rh Al2O3 750–900 Over 60 cycles of 59
reforming reaction and
the reverse (methana-
tion) carried out
(Weizmann Institute)
Worner [57] Rh Foam structures: 700–860 Both inserts gave some C 64
α-Al2O3, SiC deposition (Deutch
Zentrum Luft und
Raumfahrt)

6.4.3 Some Recent Publications on CO2 Reforming


of Methane

In conclusion, we list in Table 6.14 a number of recent publications on the subject


of CO2 reforming chosen rather at random from the large number that have been
published recently. Many of the themes that we have highlighted above return
frequently. While some of the papers report significant improvements in previously
described catalyst systems (e.g. [110, 111, 123]), others use the reaction purely as a
test reaction (generally at rather lower temperatures) to provide a method of
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 153

Table 6.14 Some recent papers (2013) on methane dry reforming


First author/ Temperature
reference Metal Support range/ C Comments
Asencios [107] Ni MgO–ZrO2 750 Coupling of partial oxidation
and dry reforming leads
to stable operation.
Optimum behaviour with
20w% Ni
Zanganeh [108] Ni MgO 550–750 Best results with catalyst
derived from NiO/MgO
solid solution with
relatively low Ni content:
Ni0.03 Mg0.97O
Zanganeh [109] Ni MgO 700 Ni0.1 Mg0.9O samples
prepared by surfactant-
assisted co-precipitation
gave good activities and
stabilities (122 h on
stream)
Chen [110] Ni CeAlO3–Al2O3 800 Presence of CeAlO3 phase
reduces graphitic carbon
deposition; this phase
decomposes CO2 giving
active surface oxygen
Xiao [111] Ni MgO (ex solid 760 Preparation based on
solution) nano-platelets giving
homogeneous
distribution of the Ni
Bhavani [111] Ni–Co NiCoMnCeZrO 800 Mixed CH4/O2/CO2 feed.
Ni–Mn and other NiCoMnCeZrO gives
combinations high activity with little
activity loss
Benrabaa [112] Ni NiFe2O4 800 Non-stoichiometric mixed
spinel formed under
reaction conditions
NixFe2-xO4 Activity improved by
reduction
Gardner [113] Ni Ba0.75NiyAl12- 900 Activity increases with
yO19-δ amount of Ni but C
deposition also increases
Baeza [114] Ni CaO–La2O3–ZrO2 400–500 CaO (2 wt%) acts as a source
of CO and active O
species that remove C
deposited on surface
Odedairo [115] Ni CeO2 700 Plasma treatment of catalyst
gives clean
metal–support interface
and higher activity and
stability than untreated
sample
(continued)
154 J.R.H. Ross

Table 6.14 (continued)


First author/ Temperature
reference Metal Support range/ C Comments
Ozkara-Aydinoglu Ni/Pt Al2O3 580–620 Amount of Pt added has
[116] major effect on the
reaction kinetics, also on
level of inhibition by CO
Tao [117] Ni Al2O3 700 Sample prepared by
single-step surfactant-
templating method more
active and stable than
conventional sample
Ma [118] Ni Carbon nanotubes 750 Samples with Ni inside
nanotubes more active
and stable than those with
Ni outside the tubes
Sun [119] Ni Mg (and or Ca) 750 NiO/MgO solid solution
O/Al2O3 formed giving active and
stable catalyst; CaO
addition causes decreased
activity and stability
Zhu [120] Ni CeO2/SiO2 800 Investigation of the effect of
calcination atmosphere.
Calcination in Ar gives
better performance
Albarazi [121] Ni CeO2/ZrO2/SBA- 600–800 Promotion of Ni within pores
15 by Ce–Zr–O improves
behaviour but gives a
decrease in accessibility
of the active sites
San Jose-Alonso Ni or Co Al2O3 700–800 Low metal contents gave low
[122] rates of C deposition.
Tests for 24 h
Nagaraja [123] Ni K–MgO/ZrO2 550–800 Presence of K+ incorporated
during the
co-precipitation of the
MgO–ZrO2 support
greatly increases the
stability. No C deposition
detected with most active
and stable samples

characterising a catalyst system (e.g. [116, 121]). One example of an improvement


is to be found in work from this laboratory ([123]; see also [87]) that shows that a
very stable catalyst with good activity can be achieved with the Ni–Zr–Mg–O
system as long as K+ ions are included in the structure.
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 155

6.5 Conclusions

The dry reforming of methane has provided a fertile area of research for more than
two decades, and some significant advances have been made in the design of
catalysts suitable for the reaction. However, it is at present unlikely that the reaction
will be used in practice unless it is combined with either partial oxidation or steam
reforming in order to avoid the problems associated with carbon formation. This
review has concentrated almost exclusively on the CO2 reforming of methane,
paying little attention to the reactions of other hydrocarbons. However, many of the
constraints encountered equally to the reforming of other molecules such as naph-
tha with the added complication that methane may be formed as a product.
Dedication. The author wishes to dedicate this review to the memory of the late
Laszlo Guczi with whom he collaborated in a joint project under the auspices of an
ERA-Chemistry project (coordinated by Alain Keinemann) until shortly before his
death in December 2012.

Acknowledgement The author wishes to thank Bhari Mallanna Nagaraja for having contributed
some of the references included in this review and for having read and commented on the concept
manuscript.

References

1. Fish JD, Hawn DC (1987) Closed-loop thermochemical energy-transport based on CO2


reforming of methane – balancing the reaction systems. J Solar Energy Eng Trans ASME
109:215–220
2. Gadalla AM, Bower B (1988) The role of catalyst support on the activity of nickel for the
reforming of methane with CO2. Chem Eng Sci 43:3049–3062
3. Gadalla AM, Somer ME (1989) Synthesis and characterization of catalysts in the system
Al2O3-MgO-NiO-Ni for methane reforming with CO2. J Am Ceram Soc 72:683–687
4. Rostrup Nielsen JR (1984) Catalytic steam reforming. In: Anderson JR, Boudart M (eds)
Catalysis – science and technology, vol 5. Springer, Berlin/Heidelberg/New York/Tokyo,
pp 1–117
5. Ross JRH (1975) The steam reforming of hydrocarbons. In: Thomas JM, Roberts MW (eds)
Surface and defect properties of solids, vol 4. The Chemical Society, London, pp 34–67
6. Rostrup-Nielsen JR (1994) Aspects of CO2 reforming of methane. In: Curry-Hyde HE, Howe
RF (eds) Natural gas conversion, II, vol 81, Studies in surface science and catalysis. Elsevier,
Amsterdam, pp 25–41
7. The Midrex Process (2013) http://www.midrex.com
8. Dibbern HC, Olesen P, Rostrup Nielsen JR, Udengaard NR (1986) Make low H2/CO syngas
using sulfur passivated reforming. Hydrocarb Process 65:71–74
9. Rostrup Nielsen JR (1984) Sulfur-passivated nickel catalysts for carbon-free steam reforming
of methane. J Catal 85:31–43
10. Ross JRH (1985) Metal catalysed methanation and steam reforming. In: Bond GC, Webb G
(eds) Catalysis, vol 7, Specialist periodical report. Royal Society of Chemistry, London,
pp 1–45
156 J.R.H. Ross

11. Keller GE, Bhasin MM (1982) Synthesis of ethylene via oxidative coupling of methane.
1 Determination of active catalysts. J Catal 73:9–19
12. Hinsen W, Baerns M (1983) Oxidative coupling of methane to C2 hydrocarbons in the
presence of different catalysts. Chemiker Zeitung 107:223–226
13. Ito T, Wang J-X, Lin C-H, Lunsford JH (1985) Oxidative dimerization of methane over a
lithium-promoted magnesium-oxide catalyst. J Am Chem Soc 107:5062–5068
14. Lunsford JH (1991) The catalytic conversion of methane to higher hydrocarbons. “Natural
gas conversion”. Stud Surf Sci Catal 61:3–13
15. Hutchings GJ, Scurrell MS, Woodhouse JR (1989) Oxidative coupling of methane using
oxide catalysts. Chem Soc Rev 18:251
16. Baerns M, Ross JRH (1992) Catalytic chemistry of methane conversion. Thomas JM,
Zamaraev KI (eds) Perspectives in catalysis. Blackwell Scientific Publishers, Oxford,
IUPAC monograph, pp 10–30
17. See also a collection of review papers in Wolf EE (eds) (1992) Methane conversion by
oxidative processes – fundamental and engineering aspects. Van Nostrand Reinhold catalysis
series, Van Nostrand Reinhold, New York
18. Holmen A, Jens K-J, Kolboe S (eds) (1990) Natural gas conversion, vol 61, Stud Surf Sci
Catal. Elsevier, Amsterdam
19. Howe R, Curry-Hyde E (eds) (1994) Natural gas conversion III, vol 81, Stud Surf Sci Catal.
Elsevier, Amsterdam
20. de Pontes M, Espinoza RL, Nicolaides C, Scholz JH, Scurrell MS (eds) (1997) Natural gas
conversion IV, vol 107, Stud Surf Sci Catal. Elsevier, Amsterdam
21. Parmaliana A, Sanfilippo D, Frusteri F, Vaccari A, Arena F (eds) (1998) Natural gas
conversion V, vol 119, Stud Surf Sci Catal. Elsevier, Amsterdam
22. Ross JRH (2005) Natural gas reforming and CO2 mitigation. Catal Today 100:151–158.
O’Connor AM (1994) BSc thesis, University of Limerick
23. Rostrup Nielsen JR (1993) Production of synthesis gas. Catal Today 18:305–324
24. Edwards JH, Maitra AM (1994) The reforming of methane with carbon dioxide – current
status and future applications. Stud Surf Sci Catal 81:291–296
25. Seshan K, Ten Barge HW, Hally W, Van Keulen ANJ, Ross JRH (1994) Carbon dioxide
reforming of methane in the presence of nickel and platinum catalysts supported on ZrO2.
Stud Surf Sci Catal 81:285–290
26. Uchijima T, Nakamura J, Sato K, Aikawa K, Kubushiro K, Kummori K (1994) Production of
synthesis gas by partial oxidation of methane and reforming of methane by carbon dioxide.
Stud Surf Sci Catal 81:325–327
27. Cant NW, Dümpelmann R, Maitra AM (1997) A comparison of nickel and rhodium catalysts
for the reforming of methane by carbon dioxide. Stud Surf Sci Catal 107:491–496
28. Slagtern A, Olsbye U, Blom R, Dahl IM (1997) The influence of rare earth oxides on
Ni/Al2O3 catalysts during CO2 reforming of CH4. Stud Surf Sci Catal 107:497–502
29. Yu C-C, Lu Y, Ding X-J, Shen S-K (1997) Studies on Ni/Al2O3 Catalyst for CO2 reforming
of CH4 to synthesis gas. Stud Surf Sci Catal 107:503–510
30. Zhang Z-L, Verykios X (1997) Performance of Ni/La2O3 catalyst in carbon dioxide
reforming of methane to synthesis gas. Stud Surf Sci Catal 107:511–516
31. Erdöhelyi A, Fodor K, Solymosi F (1997) Reaction of CH4 with CO2 and H2O over supported
Ir catalyst. Stud Surf Sci Catal 107:525–530
32. Van Keulen ANJ, Hegarty MES, Ross JRH, Van Oosterkamp PF (1997) The development of
platinum-zirconia catalysts for the CO2 reforming of methane. Stud Surf Sci Catal
107:537–546
33. Kikuchi E, Chen Y (1997) Low-temperature syngas formation by CO2 reforming of methane
in a hydrogen-permselective membrane reactor. Stud Surf Sci Catal 107:547–553
34. Ponelis AA, Van Zyl PGS (1997) CO2 reforming of methane in a membrane reactor. Stud
Surf Sci Catal 107:555–560
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 157

35. Hallische H, Bouarab R, Cherife O, Bettahar MM (1998) Effect of metal additives on


deactivation of Ni/α-Al2O3 in the CO2-reforming of methane. Stud Surf Sci Catal
119:705–710
36. Buarab R, Menad S, Hallische D, Cherifi O, Bettahar MM (1998) Reforming of methane with
carbon dioxide over supported Ni catalysts. Stud Surf Sci Catal 119:717–722
37. Nichio NN, Casella ML, Ponzi EN, Ferretti OA (1998) Ni/Al2O3 catalysts for syngas
obtention via reforming with O2 and/or CO2. Stud Surf Sci Catal 119:723–728
38. Gronchi P, Centola P, Kaddouri A, Del Rosso R (1998) Transient reactions in CO2 reforming
of methane. Stud Surf Sci Catal 119:735–740
39. Provendier H, Petit C, Estournes C, Keinnemann A (1998) Dry reforming of methane.
Interest of La-Fe-Ni solid solutions compared to LaNiO3 and LaFeO3. Stud Surf Sci Catal
119:741–746
40. Kroll VCH, Tjatjopoulos GJ, Mirodatos C (1998) Kinetics of methane reforming over
Ni/SiO2 catalysts based on a step-wise mechanistic model. Stud Surf Sci Catal 119:753–758
41. Kim J-H, Suh DJ, Park T-J, Kim K-L (1998) Improved stability of nickel-alumina aerogel
catalysts for carbon dioxide reforming of methane. Stud Surf Sci Catal 119:771–776
42. York APE, Suhartanto T, Green MLH (1998) Influence of molybdenum and tungsten dopants
on nickel catalysts for dry reforming of methane with carbon dioxide to synthesis gas. Stud
Surf Sci Catal 119:777–782
43. Suzuki S, Hayakawa T, Hamakawa S, Suzuki K, Shishido T, Takehira K (1998) Sustainable
Ni catalysts prepared by SPC method for CO2 reforming of CH4. Stud Surf Sci Catal
119:783–788
44. Stagg SM, Resasco DE (1998) Effect of promoters on supported Pt catalysts for CO2
reforming of CH4. Stud Surf Sci Catal 119:813–818
45. O’Connor AM, Meunier FC, Ross JRH (1998) A kinetic and in-situ DRIFT spectroscopy
study of carbon dioxide reforming over a Pt/ZrO2 catalyst. Stud Surf Sci Catal 119:819–824
46. Quincoces CE, Perez de Vargas S, Diaz A, Montes M, González MG (1998) Morphological
changes of Ca promoted Ni/SiO2 catalysts and carbon deposition during CO2 reforming of
methane. Stud Surf Sci Catal 119:837–842
47. Nam JW, Chae H, Lee SH, Jung H, Lee K-Y (1998) Methane dry reforming over well-
dispersed Ni catalyst prepared from perovskite-type mixed oxides. Stud Surf Sci Catal
119:843–848
48. Tomishige K, Chen Y, Yamazaki O, Himeno Y, Koganezawa Y, Fujimoto K (1998) Carbon-
free CH4-CO2 and CH4-H2O reforming catalysts – structure and mechanism. Stud Surf Sci
Catal 119:861–866
49. Fischer F, Tropsch H (1928) Brennstoff Chem 3:39
50. Bodrov IM, Apel’baum LO (1967) Kinet Katal 8:379
51. Bodrov IM, Apel’baum LO, Temkin MI. Kinet Katal 5:696
52. Takayasu O, Hirose E, Matsuda N, Matsurra I (1991) Chem Expr 6:447
53. Ashcroft AT, Cheetham AK, Green MLH, Vernon PDF (1991) Partial oxidation of methane
to synthesis gas using carbon dioxide. Nature 352:225–226; see also Vernon PDF, Green
MLH, Cheetham AK, Ashcroft AT (1992) Catal Today 13:417
54. Solymosi F, Kutsan GY, Erdohelyi A (1991) Catalytic reaction of CH4 with CO2 over
alumina-supported Pt metals. Catal Lett 11:149
55. Masai M, Kado H, Miyake A, Nishiyama S, Tsuruya S (1988) Stud Surf Sci Catal 36:67
56. Levy M, Levitan R, Rosin H, Rubin R (1993) Solar-energy storage via a closed-loop
chemical heat pipe. Solar Energy 50:179–189
57. Worner A, Tamme R (1998) CO2 reforming of methane in a solar driven volumetric receiver-
reactor. Catal Today 46:165–174
58. Wang SB, Lu GQM, Millar GJ (1996) Carbon dioxide reforming of methane to produce
synthesis gas over metal-supported catalysts: state of the art. Energy Fuels 10:896–904
59. Hu YH, Ruckenstein E (2004) Catalytic conversion of methane to synthesis gas by partial
oxidation and CO2 reforming. Adv Catal 48:297–345
158 J.R.H. Ross

60. Ross JRH, Van Keulen ANJ, Hegarty MES, Seshan K (1996) The catalytic conversion of
natural gas to useful products. Catal Today 30:193–199
61. Ross JRH (2005) Natural gas reforming and CO2 mitigation. Catal Today 100:151–158
62. Bradford MCJ, Vannice MA (1999) CO2 reforming of CH4. Catal Rev Sci Eng 41:1–42
63. Verykios XE (2003) Catalytic reforming of natural gas for the production of chemicals and
hydrogen. Int J Hydrogen Energy 28:1045–1063
64. Armor JN (1999) The multiple roles of catalysis in the production of H2. Appl Catal A Gen
176:159–176
65. Ma J, Sun N, Zhang X (2009) A short review of catalysis for CO2 conversion. Catal Today
148:221–231
66. Kodama T (2003) High-temperature solar chemistry for converting solar heat to chemical
fuels. Progr Energy Combust Sci 29:567–597
67. Horiuchi T, Sakuma K, Fukui T, Kubo Y, Osaki T, Mori T (1996) Suppression of carbon
deposition in the CO2 reforming of CH4 by adding basic metal oxides to a Ni/Al2O3 catalyst.
Appl Catal A Gen 144:111–120
68. Wang SB, Lu GQ (1998) Role of CeO2 in Ni/CeO2-Al2O3 catalysts for carbon dioxide
reforming of methane. Appl Catal B Environ 19:267–277
69. Wang SB, Lu GQM (1998) CO2 reforming of methane on Ni catalysts: effects of the support
phase and preparation technique. Appl Catal B Environ 16:269–277
70. Kim JH, Suh DJ, Park TJ, Kim KL (2000) Effect of metal particle size on coking during CO2
reforming of CH4 over Ni-alumina aerogel particles. Appl Catal A Gen 197:191–200
71. Martinez R, Romero E, Guimon C, Bilbao R (2004) CO2 reforming of methane over
coprecipitated Ni-Al catalysts modified with lanthanum. Appl Catal A Gen 274:139–149
72. Juan-Juan J, Roman-Martinez MC, Illan-Gomez MJ (2004) Catalytic activity and character-
ization of Ni/Al2O3 and NiK/Al2O3 catalysts for CO2 methane reforming. Appl Catal A Gen
264:169–174
73. Laosiripojana N, Sutthisripok W, Assabumrungrat S (2005) Synthesis gas production from
dry reforming of methane over CeO2 doped Ni/Al2O3: influence of the doping of ceria on the
resistance toward carbon formation. Chem Eng J 112:13–22
74. Goula MA, Lemonidou AA, Efstathiou AM (1996) Characterization of carbonaceous species
formed during reforming of CH4 with CO2 over Ni/CaO-Al2O3 catalysts studied by various
transient techniques. J Catal 161:626–640
75. Osaki T, Mori T (2001) Role of potassium in carbon-free CO2 reforming of methane on
K-promoted Ni/Al2O3 catalysts. J Catal 204:89–97
76. Rostrup Nielsen JR, Hansen JHB (1993) CO2 reforming of methane over transition metals.
J Catal 144:38–49
77. Tomishige K, Chen YG, Fujimoto K (1999) Studies on carbon deposition in CO2 reforming
of CH4 over nickel-magnesia solid solution catalysts. J Catal 181:91–103. See also:
Tomishige K, Yamazaki O, Chen YG, Yokoyama K, Li XH, Fujimoto K (1998) Development
of ultra-stable Ni catalysts for CO2 reforming of methane. Catal Today 45:35–39
78. Wei JM, Iglesia E (2004) Isotopic and kinetic assessment of the mechanism of reactions of
CH4 with CO2 or H2O to form synthesis gas and carbon on nickel catalysts. J Catal
224:370–383
79. Ruckenstein E, Hu YH (1995) Carbon dioxide reforming of methane over nickel alkaline
earth metal oxide catalysts. Appl Catal A Gen 133:149–161
80. Ruckenstein E, Hu YH (1998) Combination of CO2 reforming and partial oxidation of
methane over NiO/MgO solid solution catalysts. Ind Eng Chem Res 37:1744–1747
81. Djaida A, Libs S, Keinnemann A, Barama A (2006) Characterization and activity in dry
reforming of methane on NiMg/Al and Ni/MgO catalysts. Catal Today 113:194–200
82. Guczi L, Stefler G, Geszti O, Sajo I, Paszti Z, Tompos A, Schay Z (2010) Methane dry
reforming with CO2: a study on surface carbon species. Appl Catal A Gen 375:236–246
83. Wang H, Miller JT, Shakouri M, Xi C, Wu T, Zhao H, Cem Akatay M (2013) XANES and
EXAFS studies on metal nanoparticle growth and bimetallic interaction of Ni-based catalysts
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 159

for CO2 reforming of CH4. Catal Today 207:3–12. See also: Zhang J, Wang H, Dalai AK
(2007) J Catal 249:298-; idem (2008) Appl Catal A Gen 339:321
84. Roh HS, Jun KW, Dong WS, Chang JS, Park SE, Joe YI (2002) Highly active and stable
Ni/Ce-ZrO2 catalyst for H2 production from methane. J Mol Catal A Chem 181:137–142; see
also: Li XS, Chang JS, Tian MY, Park SE (2001) CO2 reforming of methane over modified
Ni/ZrO2 catalysts. Appl Organometal Chem 15:109–112; and Li X, Chang JS, Park SE (1999)
Carbon as an intermediate during the carbon dioxide reforming of methane over zirconia-
supported high nickel loading catalysts. Chem Lett 1099–1100
85. Montoya JA, Romero-Pascaul E, Gimon C, Del Angel P, Monzon A (2000) Methane
reforming with CO2 over Ni/ZrO2-CeO2 catalysts. Catal Today 63:71–85
86. Lercher JA, Bitter JH, Hally W, Niessen W, Seshan K (1996) Design of stable catalysts for
methane-carbon dioxide reforming. Stud Surf Sci Catal 101:463–472; see also: Hally W,
Bitter JH, Seshan K, Lercher JA, Ross JRH (1994) Problem of coke formation on Ni/ZrO2
catalysts during the carbon dioxide reforming of methane. Stud Surf Sci Catal 88:167–173;
and Ref. [28]
87. Nagaraja BM, Bulushev DA, Beloshapkin S, Ross JRH (2011) The effect of potassium on the
activity and stability of Ni-MgO-ZrO2 catalysts for the dry reforming of methane to give
synthesis gas. Catal Today 178:132–136
88. Bradford MCJ, Vannice MA (1996) Catalytic reforming of methane with carbon dioxide over
nickel catalysts. Part 1. Catalyst characterization and activity; and Part 2. Reaction. Appl
Catal A Gen 142:73–96 and 97–122
89. Zhang ZL, Verykios XE, MacDonald SM, Affrossman S (1996) Comparative study of carbon
dioxide reforming of methane to synthesis gas over Ni/La2O3 and conventional nickel-based
catalysts. J Phys Chem 100:744–754; see also Zhang Z, Verykios XE (1977) Carbon dioxide
reforming of methane to synthesis gas over Ni/La2O3 catalysts. Appl Catal A Gen
138:109–133; and Verykios XE (2003) Catalytic reforming of natural gas for the production
of chemicals and hydrogen. Int J Hydrogen Energy 28:1045–1063
90. Swaan HM, Kroll VCH, Martin GA, Mirodatos C (1994) Deactivation of supported nickel
catalysts during the reforming of methane by carbon dioxide. Catal Today 21:571–578; see
also: Kroll VCH, Swaan HM, Mirodatos C (1996) Methane reforming reaction with carbon
dioxide over Ni/SiO2 catalyst. 1. Deactivation studies. J Catal 161:409–422
91. Valderrama G, Kiennemann A, Goldwasser MR (2010) La-Sr-Ni-Co-O based perovskite-
type solid solutions as catalyst precursors in the CO2 reforming of methane. J Power Sources
195:1765–1771
92. Ashcroft AT, Cheetham AK, Green MLH, Vernon PDF (1991) Partial oxidation of methane
to synthesis gas using carbon dioxide. Nature 352:225–226; see also Vernon PDF, Green
MLH, Cheetham AK, Ashcroft AT (1992) Partial oxidation of methane to synthesis gas, and
carbon dioxide as an oxidizing agent for methane conversion. Catal Today 13:417–426
93. Stagg-Williams SM, Noronha FB, Fendley G, Resasco DE (2000) CO2 reforming of CH4
over Pt/ZrO2 catalysts promoted with La and Ce oxides. J Catal 194:240–249
94. Damyanova S, Bueno JMC (2003) Effect of CeO2 loading on the surface and catalytic
behaviors of CeO2-Al2O3-supported Pt catalysts. Appl Catal A Gen 253:135–150
95. O’Connor AM, Ross JRH (1998) The effect of O2 addition on the carbon dioxide reforming
of methane over Pt/ZrO2 catalysts. Catal Today 46:203–210
96. Bitter JH, Seshan K, Lercher JA (1997) The state of zirconia supported Pt catalyst for CO2/
CH4 reforming. J Catal 171:279–286
97. Nakamura J, Aikawa K, Sato K, Uchima T (1994) Role of support in reforming of CH4 with
CO2 over Rh catalysts. Catal Lett 25:265–270
98. Richardson JT, Garrait M, Hung JK (2003) Carbon dioxide reforming with Rh and Pt-Re
catalysts dispersed on ceramic foam supports. Appl Catal A Gen 255:69–82
99. Mark MF, Maier WF (1996) CO2 reforming of methane on supported Rh and Ir catalysts.
J Catal 164:122–130
160 J.R.H. Ross

100. Erdohelyi A, Cserenyi J, Solymosi F (1993) Activation of CH4 and its reaction with CO2 over
supported Rh catalysts. J Catal 141:287–299
101. Qin D, Lapszewicz J (1994) Study of mixed steam and CO2 reforming of CH4 to syngas on
MgO-supported metals. Catal Today 21:551–560
102. Maestri M, Vlachos DG, Beretta A, Groppi G, Tronconi E (2008) Steam and dry reforming of
methane on Rh: microkinetic analysis and hierarchy of kinetic models. J Catal 259:211–222
103. Erdohelyi A, Cserenyi J, Papp E, Solymosi F (1994) Catalytic reaction of methane with
carbon dioxide over supported palladium. Appl Catal A Gen 108:205–219
104. Donazzi A, Beretta A, Groppi G, Forzatti P (2008) Catalytic partial oxidation of methane
over a 4 % Rh/alpha-alumina catalyst. Part II. Role of CO2 reforming. J Catal 255:259–268
105. Claridge JB, York APE, Brungs AJ, Marquez-Alvarez C, Sloan J, Tsang SC, Green MLH
(1998) New catalysts for the conversion of methane to synthesis gas: molybdenum and
tungsten carbide. J Catal 180:85–100 (See also reference [25])
106. Treacy D, Ross JRH (2004) Carbon dioxide reforming of methane over supported molybde-
num carbide catalysts. In: Bao X, Xu Y (eds) Natural gas conversion VII, vol 147, Studies in
surface science and catalysis. Elsevier, Amsterdam, pp 193–198
107. Asencios YJO, Assaf EM (2013) Combination of dry reforming and partial oxidation of
methane on NiO-MgO-ZrO2 catalyst: effect of nickel content. Fuel Proc Technol
106:247–252
108. Zanganeh R, Rezaei M, Zamaniyan A (2013) Dry reforming of methane to synthesis gas on
NiO-MgO nanocrystalline solid solution catalysts. Int J Hydrogen Energy 38:3012–3018
109. Zanganeh R, Rezaei M, Zamaniyan A, Bozorgzadeh HR (2013) Preparation of Ni0.1 Mg0.9O
nanocrystalline powder and its catalytic performance in methane reforming with carbon
dioxide. J Ind Eng Chem 19:234–239
110. Chen W, Zhao G, Xue Q, Chen L, Lu Y (2013) High carbon-resistance Ni/CeAlO3-Al2O3
catalyst for CH4/CO2 reforming. Appl Catal B Environ 136–137:260–268
111. Bhavani AG, Kim WY, Kim JY, Lee JS (2013) Improved activity and coke resistance by
promoters of nanosized trimetallic catalysts for autothermal carbon dioxide reforming of
methane. Appl Catal A Gen 450:63–72
112. Benrabaa R, Löfberg A, Rubbens A, Bordes-Richard E, Vannier RN, Barama A (2013)
Structure, reactivity and catalytic properties of nanoparticles of nickel ferrite in the dry
reforming of methane. Catal Today 203:188–195
113. Gardner TH, Spivey JJ, Kugler EL, Pakhare D (2013) CH4–CO2 reforming over
Ni-substituted barium hexaaluminate catalysts. Appl Catal A Gen 455:129–136
114. Baeza BB, Pedrero CM, Soria MA, Ruiz AG, Rodemerck U, Ramos IR (2013) Transient
studies of low-temperature dry reforming of methane over Ni-CaO/ZrO2-La2O3. Appl Catal
B Environ 129:450–459
115. Odedairo T, Chen J, Zhu Z (2013) Metal–support interface of a novel Ni–CeO2 catalyst for
dry reforming of methane. Catal Commun 31:25–31
116. Ozkara-Aydinoglu S, Aksoylu AE (2013) A comparative study on the kinetics of carbon
dioxide reforming of methane over Pt–Ni/Al2O3 catalyst: effect of Pt/Ni Ratio. Chem Eng J
215–216:542–549
117. Tao K, Shi L, Ma Q, Wang D, Zeng C, Kong C, Wu M, Chen L, Zhou S, Hu Y, Tsubaki N
(2013) Methane reforming with carbon dioxide over mesoporous nickel–alumina composite
catalyst. Chem Eng J 221:25–31
118. Ma Q, Wang D, Wu M, Zhao T, Yoneyama Y, Tsubaki N (2013) Effect of catalytic site
position: nickel nanocatalyst selectively loaded inside or outside carbon nanotubes for
methane dry reforming. Fuel 108:430–438
119. Sun Y, Collins M, French D, McEvoy S, Hart G, Stein W (2013) Investigation into the
mechanism of NiMg(Ca)bAlcOx catalytic activity for production of solarised syngas from
carbon dioxide reforming of methane. Fuel 105:551–558
6 Syngas Production Using Carbon Dioxide Reforming: Fundamentals and Perspectives 161

120. Zhu J, Peng X, Yao L, Deng X, Dong H, Tong D, Hu C (2013) Synthesis gas production from
CO2 reforming of methane over NieCe/SiO2 catalyst: the effect of calcination ambience. Int J
Hydrogen Energy 38:117–126
121. Albarazi A, Beaunier P, Costa PD (2013) Hydrogen and syngas production by methane dry
reforming on SBA-15 supported nickel catalysts: on the effect of promotion by
Ce0.75Zr0.25O2 mixed oxide. J Hydrogen Energy 38:127–139
122. San Jose-Alonso D, Illan-Gomez MJ, Roman-Mart{nez MC (2013) Low metal content Co
and Ni alumina supported catalysts for the CO2 reforming of methane. J Hydrogen Energy
38:2230–2239
123. Nagaraja BM, Bulushev DA, Belshapkin S, Chansai S, Ross JRH (2013) Potassium-doped
Ni–MgO–ZrO2 catalysts for dry reforming of methane to synthesis gas. Topics Catal
56:1686–1694; see also Ref. [87]
Chapter 7
Carbon Dioxide as C-1 Block
for the Synthesis of Polycarbonates

Peter T. Altenbuchner, Stefan Kissling, and Bernhard Rieger

Abbreviations

BDI β-diiminate
BO 1-butene-oxide
CHC cyclohexene carbonate
CHO cyclohexene oxide
DFT density functional theory
DMAP Dimethylaminopyridine
DNP 2,4-dinitrophenolate
OAc acetate
PC propylene carbonate
PDI Polydispersity index
PCHC poly(cyclohexene carbonate)
PO propylene oxide
PPC poly(propylene carbonate)
PPNCl bis(triphenylphosphine)iminium chloride
TBD 1,5,7-triabicyclo[4,4,0]dec-5-ene
Tg glass transition temperature
Tm melting point
TMS Trimethylsilyl-
TOF turn-over frequency
TON turn-over number
VCHC vinylcyclohexene oxide

P.T. Altenbuchner • S. Kissling • B. Rieger (*)


Technische Universität München, WACKER-Lehrstuhl für Makromolekulare Chemie,
Lichtenbergstrasse 4, Garching 85748, Germany
e-mail: rieger@tum.de

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 163
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_7,
© Springer-Verlag Berlin Heidelberg 2014
164 P.T. Altenbuchner et al.

ZnAA zinc-adipate
ZnGA zinc-glutarate
ZnPA zinc-pimelate
ZnSA zinc-succinate

7.1 Introduction

Fossil energy sources became the dominant energy carriers at the beginning of the
twentieth century. Since then, no alternatives to crude oil, gas, and coal have arisen to
supplant them as a universal source of energy and raw materials for the chemical
industry [1, 2]. The discovery of versatile possibilities for the use of crude oil as well
as its low price and historical abundance made the steep rise in living standards in
the last 100 years possible. As a result, the concentration of carbon dioxide in the
atmosphere has risen significantly and reached 397 ppm in 2013 [3]. However, the
finiteness of coal, gas, and especially oil reserves and the steadily increasing demand
for them are likely to create a rising imbalance of supply and demand. Furthermore,
the various negative implications of reliance on fossil energy carriers generate only
limited incentives for industry and consumers to change their accustomed behavior.
The reason for the disproportionate low price of CO2 is that the resulting environ-
mental damage is not or is only partially included in it. The rising generation of
carbon dioxide however opens up a vast C-1 source for the synthesis of chemical raw
materials. The attractive properties of CO2 as a nontoxic, renewable, and low-cost
C1-building block are known, and it is currently used on an industrial scale in the
production of urea (146  106 t/a), methanol (6  106 t/a), cyclic carbonate
(0.040  106 t/a), and salicylic acid (0.060  106 t/a), among others [4]. However,
the thermodynamic stability has thus far hampered its usage as widespread chemical
reagent. Methods to overcome the high energy barriers are based upon reduction,
oxidative coupling with unsaturated compounds on low valency metal complexes,
and increasing electrophilicity of the carbonyl carbon [5].
The reaction of CO2 with highly reactive substrates, such as epoxides, affords
materials which can be produced on industrially relevant scale (Fig. 7.1). These
polycarbonates from CO2 have the potential to be the next industrial product with a
high impact on the usage as commodity polymers. Consequently, the synthesis of
highly active catalysts and understanding their mechanistic behavior is of upmost
importance and the current developments will be discussed in this chapter.

7.2 Current Situation

Over 20 chemical reaction pathways for carbon dioxide have been developed
over the course of the last two decades. Industrially viable reactions are scarce
and only the production of urea, salicylic acid, and carbonates have reached
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 165

Fig. 7.1 Alternating


copolymerization of
propylene oxide (PO) (a)
and cyclohexene oxide
(CHO) (b) with CO2

larger volumes [6, 7]. The amount of carbon dioxide consumed by these or any
other chemical processes however cannot counteract anthropologically caused
climate change. They rather have the potential to supplant oil-based starting
materials in various synthetic pathways and thereby open up a more eco-friendly
synthesis. The production of biodegradable polycarbonates from CO2 and epoxides
is currently receiving a lot of attention [5, 8, 9]. Aside from the polymeric product,
cyclic carbonates also have potential applications such as in lithium ion batteries as
electrolytes and as intermediates for the synthesis of linear dialkyl carbonates [4].
In this chapter cyclic carbonates will not be discussed in detail [9–11]. The focus
will instead be on the catalysis behind the synthesis of polymeric materials with
the recent developments at the center. Polycarbonates such as poly(propylene
carbonate) (PPC) and poly(cyclohexene carbonate) (PCHC) have promising phys-
ical properties like lightness, durability, biodegradability, heat resistance, high
transparency, and gas permeability which make them possible candidates for
applications in the automotive, medical, and electronics industries.

7.2.1 Properties of Polycarbonates on the Example


of Poly(propylene carbonate)

A large variety of epoxides are commercially available and subsequently have


been tested for their behavior in copolymerization with carbon dioxide or even in
terpolymerization with an additional epoxide. In industry though only ethylene
oxide and propylene oxide are currently produced on large scale and have
therefore a commercial advantage. Terpolymerizations promise a synthetic
approach toward materials with better physical and material properties. Poly
(propylene carbonate) and poly(cyclohexene carbonate) do not yet meet the
demands of industry in respect of their glass transition temperatures, achieved
molecular weights, elasticity, and their homogeneity. This justifies an academic
approach toward better understanding of the polymerization behavior and
properties of the produced materials.
166 P.T. Altenbuchner et al.

Fig. 7.2 Poly(propylene carbonate) (left), poly(cyclohexene carbonate) (right)

7.2.1.1 General Properties

Poly(propylene carbonate) is an aliphatic polycarbonate (Fig. 7.2). Due to irregular


incorporation of carbon dioxide, polyether units can be presented in the polymer
backbone. PPC is amorphous, hydrophobic, and soluble in organic solvents, e.g.,
dichloromethane, acetone, and cyclic propylene carbonate. The general properties
strongly depend on the molecular weight, polydispersity, ratio of carbonate to ether
linkages, microstructure, and possible catalyst contaminations in the polymer
[12, 13]. Due to all the mentioned factors in literature, a temperature range from
25 to 46  C for the glass transition temperature of PPC is given, and also other
material characteristics like tensile strength and Young’s modulus also show some
variance [14, 15]. Cyclic carbonates act as plasticizers in the polymer and have the
same effect as ether units in the polymer backbone: they decrease the glass
transition temperature. Additionally, high amounts of ether linkages affect the
material’s thermal stability.
As the available analytical data is not always sufficient to correctly compare the
investigated polymer samples, some properties only can be given as approximate
ranges. For polymer samples with a Tg of about 40  C, the range for the modulus
was from 700 to 1,400 MPa, whereas for PPC with a Tg of 30  C the modulus was
reported to be 200 and 1,000 MPa [14, 15]. However, for strictly alternating PPC
with less than 1 % cyclic carbonate content, an elongation at break of less than 50 %
and a tensile strength of around 40 MPa (Mn ¼ 180,000 g/mol) is reported in the
literature [8]. One of the major disadvantages of PPC is its tendency to degrade at
elevated temperatures. Earlier reports tested PPC with different molecular weights
at temperatures between 150 and 180  C. They observed severe degradation which
of course has negative consequences for the material properties. Current publica-
tions suggest that the degradation is not solely induced by temperature, but that
residual catalyst can enhance the degradation as well [16–20]. Already at 150  C, a
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 167

PPC sample with a molecular mass over 200 kg/mol and a PDI of 1.2 show severe
degradation to cyclic carbonate after 1 h if the catalyst residue is not completely
removed from the polymer [12]. An increase in thermostability to 180  C was made
possible by a thorough catalyst removal from the sample. Further increases in
temperature resulted in profound degradation even for the treated polymer sample.
Blending of PPC with other engineering plastics is possible but difficult due to the
fact that the optimum melt processing temperature for PPC is around 100–140  C
[21]. In general, polyesters and polyamides exhibit higher melting temperatures of
above 200  C [16]. The catalysts known in the literature so far show low produc-
tivity and produce at most 38 kg polymer per gram cobalt catalyst [22]. The
catalysts cannot be left in the polymer due to their coloring- and degradation-
enhancing properties. Furthermore, regulations concerning the amount of toxic
metals like cobalt make a thorough clean up of the polycarbonates necessary.

7.2.1.2 Degradation Mechanism and Thermal Stability

In the previous subchapter, the thermal degradation of PPC was briefly addressed.
Now the underlying mechanisms and possibilities for the formation of more stable
PPC shall be discussed in more detail.
One form of thermal degradation is chain unzipping, whereby the nucleophile
for the degradation can be located in the polymer chain or through an external
nucleophile from another chain (Fig. 7.3a). Terminal hydroxyl groups can perform
intramolecular backbiting reactions and thereby cause the formation of cyclic
carbonates. The backbiting reaction can occur via a terminal alkoxide chain end
which attacks at a carbonyl carbon. The backbiting route can also be via a terminal
carbonate which attacks at an electrophilic carbon in the polymer chain.
Intermolecular back biting leads to an attack at a random position in the polymer
and forms shorter chains. The second degradation mechanism, which is assumed to
take place increasingly at higher temperatures, is random chain scission (Fig. 7.3b)
[23, 24]. At elevated temperatures, a decarboxylation reaction takes place and gives
rise to carbon dioxide and an olefin moiety. As can be seen in Fig. 7.3c, catalyst
impurities, water residues, acids, bases, and solvents promote degradation reactions
and should therefore be thoroughly excluded or removed upon completion of the
reaction [16].
The thermal stabilities reported in the literature are difficult to compare and to
relate to one another due to a lack of information concerning both the
polycarbonates and the testing conditions. Generally, the chain unzipping of poly
(propylene carbonate) and other carbonates can be suppressed to a certain degree by
end-capping, although the potential for depolymerization via chain scission is not
affected by this alteration. In the end-capping process, the nucleophilic terminal
hydroxyl groups of the poly(alkylene carbonates) are replaced with less reactive
functionalities (e.g., isocyanates, carboxylic acid anhydrates, acetates, urethanes,
sulfonates, phosphorous containing compounds) [23–27]. Analytical data
concerning the remaining hydroxyl groups have not been published so far.
168 P.T. Altenbuchner et al.

Fig. 7.3 Backbiting mechanisms for poly(propylene carbonate): (a) chain unzipping, (b) chain
scission, and (c) residual water traces promoting the degradation

Hydrogen bonding has also been shown to improve the thermal stability of
polycarbonates significantly. In a recent publication, octadecanoic acid was added
to PPC by solution mixing. The resulting supramolecular complexes showed
improved thermal stability, a thermotropic liquid crystalline character, and
enhanced glass transition compared to an amorphous PPC copolymer [28]. Improve-
ments in thermal stability were also obtained by adding wood flour [29] or
nanoparticles [30–33] to polycarbonates by blending with other polymers [34–37].
The actual measured thermal stabilities are difficult to compare due to different
testing conditions and the lack of information concerning the poly(propylene carbon-
ates) themselves. It was reported that interactions between poly(propylene carbonate)
and metal ions can in fact enhance the thermal stability [38–40]. Nevertheless, none of
the modifications to poly(propylene carbonate) mentioned in the literature are able to
inhibit degradation during processing at temperatures over 180  C.

7.2.1.3 Terpolymerization

Terpolymerizations and block copolymerizations have the potential to overcome the


poor properties of poly(propylene carbonate) and poly(cyclohexene oxide), respec-
tively, and instead combine their strengths in one polymeric material. Other epoxides
such as styrene oxide and epichlorohydrin offer additional opportunities as copoly-
mers and also as substitutes for PO or CHO. Consequently, both the copolymerization
and the terpolymerization of carbon dioxide and various epoxides have been inves-
tigated (Fig. 7.4) [41–54]. The incorporation of other monomers (lactones and
anhydrides) in polycarbonate structures has also been studied [55–62].
Fundamental research on the reasons for the differing reactivities and on the
inactivity of some catalysts toward some monomers is scarce. Rieger et al. published
detailed kinetic studies supported by quantum chemical calculations in 2011 [63]. The
homogeneous dinuclear zinc catalyst 26 of Williams et al. was chosen as a test system
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 169

Fig. 7.4 Some epoxides used in the literature [41–54]

as it exhibits the highest activities for CHO/CO2 copolymerization at one bar carbon
dioxide pressure (TOF ¼ 25 h1) [64]. The copolymerization of propylene oxide and
carbon dioxide however is not possible with catalyst 26: instead, only cyclic carbonate
is formed. Many zinc-based catalyst systems share this inactivity for PO/CO2 copo-
lymerization or possess only low activities compared to other metals. Therefore,
Rieger et al. systematically varied the reaction conditions to identify the underlying
principles. They found that neither propylene oxide nor additionally added cyclic
propylene carbonate does deactivate the catalyst for the formation of PCHC, although
in the presence of cyclic propylene carbonate, the activity was decreased dramatically.
Terpolymerizations with different ratios of CHO to PO resulted only in PCHC and no
incorporation of PO was found, which can be attributed to a very fast backbiting
reaction in the case of PO. DFT calculations showed that the ring strains of CHO and
PO are comparable and that kinetic barriers to polymer formation in theory favor
generation of PPC over PCHC. Consequently, the authors concluded that in the case of
the studied zinc catalyst, the depolymerization rate of PPC is several magnitudes faster
than that of PCHC. The limitations of zinc complexes gave cobalt catalysts an
advantage in the field of CHO/PO and CO2 terpolymerizations. In 2006 the first
(salen)CoX catalysts with increased activities were discovered [46]. Catalyst 1, in
combination with [PPN]Cl, is able to terpolymerize at 25  C and 15 bar CO2 pressure
equimolar quantities of PO and CHO with activities up to 129 h1. The resulting
polymer had a narrow PDI of 1.24 and a high content of carbonate linkages and
showed a single Tg. The authors varied the reaction conditions and were able to tune
the CHO content from 30 to 60 % which gave polymers with Tg in the range of
50–100  C. The highly alternating nature of the polymer was attributed to the
inhibiting nature of CHO on PO which leaves both epoxides with matched reactivities
in the copolymerization. The regioselective ring opening of PO was not disturbed by
the competing binding of the epoxides to the metal center. Another cobalt-based
catalyst (7) with covalently bound ionic groups was published in 2010 by Lee
et al. They performed a multitude of experiments with various epoxides in different
combinations [65]. Terpolymerizations of CO2/PO/CHO, CO2/PO/1-hexene oxide
(HO), and CO2/PO/1-butene oxide (BO) were carried out. Activities ranged between
4,400 and 14,000 h1 and no cyclic carbonate formation was observed during the
reactions. Furthermore, an increase in decomposition temperature was achieved by
employing a third monomer and the results demonstrated that the Tg of PO/CO2 can be
170 P.T. Altenbuchner et al.

Fig. 7.5 Cobalt catalysts for the co- and terpolymerization of carbon dioxide and epoxides [22,
46, 65–67]

adjusted at will in a certain range (0–100  C) by using a third monomer (CHO, HO, or
BO). A linear dependency between the glass transition temperature in the terpolymer
and the proportions of the third monomer was observed, which makes Tg adjustments
possible. For the terpolymer CO2/PO/HO, the temperature range was found to be from
15 to 32  C and for CO2/PO/BO from 9 to 33  C. The authors also report that the
incorporation of further epoxides (styrene oxide (SO), isobutylene oxide (IO), and
glycidyl ether (GE)) in the PO/CO2 copolymer failed under the tested reaction
conditions due to remaining impurities in the monomers. Almost at the same time
Lu et al. reported a bifunctional cobalt(III) catalyst (3) for the copolymerization of
CHO/CO2 and for terpolymerization with additional aliphatic epoxides such as PO,
EO, BO, and HO (Fig. 7.5) [66].
The catalyst showed high activities (TOF) of 1,958–3,560 h1 (90  C, 25 bar, 1:
1 ¼ CHO: [PO,BO,HO,EO]) and molecular weights of the resulting terpolymers
ranged between 40,000 and 60,000 g/mol (PDI ~ 1.1). The produced polymers had
a content of 37–65 % of cyclohexene carbonate linkages and only one Tg was
measured (32–79  C).
Functionalized epoxides open up the possibility for post modification of the resulting
carbonate structures. Already in 2007 Coates et al. used β-diiminate zinc(II) acetate
catalysts ([(BDI)ZnOAc]2) for the terpolymerization of CHO, vinylcyclohexene oxide
(VCHO), and carbon dioxide to synthesize vinyl-functionalized polycarbonates
[43]. The subsequent olefin metathesis using a Grubbs’ catalyst [68] made the transfor-
mation of linear polycarbonates into nanoparticles of controlled size possible. The
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 171

Fig. 7.6 BDI zinc catalyst


with a norbornene
carboxylate initiator [69]

employed epoxides had comparable reactivities, which enabled the generation of


terpolymers with a VCHO content of 30 mol%. Medium molecular weights of
54,100 g/mol and low PDI (Mw/Mn ¼ 1.20) were reached. It was observed that the
cross-metathesis reaction progressed quickly in the beginning, but due to the formation
of increasingly rigid polymer chains slowed down after 15 min. A remarkable increase
in the Tg was observed from 114 to 194  C at 76 % cross-linking. This increase was
attributed to reduced chain mobility. A similar approach was used by Coates et al. in
2012 for the synthesis of multisegmented graft polycarbonates [69]. Norbornenyl-
terminated multiblock poly(cyclohexene carbonate)s were again synthesized using a
BDI zinc catalyst 8 with a norbornene carboxylate initiator (Fig. 7.6). The living nature
of the copolymerization allowed the step-by-step addition of functionalized CHO. This
generated variable block sequences which could subsequently be transformed to seg-
mented graft copolymers by ring-opening metathesis polymerization.
Propylene oxide derivatives in terpolymers can also be found in the literature.
Frey et al. terpolymerized 1,2-isopropylidene glyceryl glycidyl ether (IGG),
glycidyl methyl ether (GME), and carbon dioxide [70]. After the acidic removal
of the protecting acetal groups, a polycarbonate poly((glyceryl glycerol)-co-
(glycidyl methyl ether) carbonate) (P((GG-co-GME)C)) is received. The analysis
of the polymeric material showed monomodal molecular weight distributions and
PDIs between 2.5 and 3.3, although the molecular mass was comparably low
(12,000–25,000 g/mol). The Tg of the unprotected polymer was 3–5  C higher
than those of the corresponding protected copolymers. Interestingly, during inves-
tigations of the degradation behavior of the P((GG-co-GME)C) copolymers in THF
solution, no backbone degradation was observed over 21 days. The authors ascribe
this observation to the reduced stability of the resulting cyclic carbonate structures.
As can be seen in this chapter, the possibilities for further research are far from
exhausted. Apart from the synthesis of ever more refined structures, the analytics
must evolve simultaneously to access the microstructure. Therefore, current devel-
opments in the field of polymer NMR techniques will be discussed in detail in the
next chapter.
172 P.T. Altenbuchner et al.

Fig. 7.7 The four proposed regiosequences with central HT junctions

7.2.1.4 Microstructure of Poly(propylene carbonate)

Since its first discovery, poly(propylene carbonate) has been characterized via a
multiple analytical methods by various research groups [71–76]. NMR has been
used extensively for microstructure analysis because PPC with higher stereoregular-
ity has the potential to increase the glass transition temperature (Tg). In general, the
ring opening of PO occurs via an SN2 type mechanism at the methylene position
resulting in a predominantly head-to-tail (HT) regiostructure of the CO2/PO copol-
ymer. By contrast, tail-to-tail (TT) or head-to-head (HH) connections arise if the ring
opening of the epoxide takes place at the methine carbon. In 2002, Chisholm and
colleagues assigned 13C NMR spectra for different PPC copolymers using a statistical
approach [17]. First, they studied the 13C NMR spectrum of (S)-PPC obtained from
the copolymerization of (S)-PO and CO2 using a zinc glutarate catalyst. The carbon-
ate carbon region showed more than one resonance, providing information about the
possible regiosequences that occur at the tetrad level. Consequently, Chisholm
et al. proposed four regiosequences with central head-to-tail (HT) junctions that are
distinguishable by NMR for the carbonate carbons (Fig. 7.7).
The other carbonate carbon signals arising from TT and HH junctions at the diad
level were assigned as s. Chisholm and colleagues applied the same statistical
approach at the triad level for aliphatic carbon in PPC and later on they analyzed
the structure of oligoether carbonates [17, 77]. Those compounds were potential
models for the PPC microstructural assignments in NMR studies. It was possible to
show with 13C NMR investigations that the carbonate carbon signals have both
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 173

Fig. 7.8 13C (125 MHz, CDCl3) NMR spectra (carbonate carbon region): (a) (R)-PPC copolymer,
(b) (R)-PPC copolymer synthesized, (c) (S)-PPC copolymer, and (d) (S)-PPC copolymer, 1 equiv-
alent [PPN]Cl, 1,000 eq. PO at 30  C and 30 bar CO2 for 20 h [(a) and (d) ((R,R)-(salen)CoCl) (b)
and (c) (S,S)-(salen)CoCl] (Reprinted with permission from Ref. [78]. Copyright 2012, American
Chemical Society)

regio- and stereosensitivity at the diad and tetrad levels to their adjacent ether units
[77]. Density functional theory calculations were applied to the tested oligoether
carbonates. These calculations indicated that the carbonate groups exist predomi-
nantly in cis-cis geometries with more than one stable conformation for each
molecule. In addition, the 13C chemical shifts predicted in the calculations were
sensitive to the conformations of the molecules and the configurations of the
stereocenters in PO ring-opened units. Recently, different poly(propylene carbon-
ate) (PPC) microstructures have been synthesized by Rieger et al. from the alter-
nating copolymerization of CO2 with propylene oxide (PO) using chiral cobalt and
(salen)Cr catalysts [78, 79].
The model which was presented in the study of Rieger et al. was based on triad
structure obtained from 13C NMR spectroscopy (HT carbonyl region). The 13C
NMR spectra of selected poly(propylene carbonate) samples were recorded using a
900 MHz (1H) spectrometer, showing a previously unreported fine splitting of the
carbonate resonances. Utilization of enantiopure and enantio-enriched PO allowed
to apply the approach of isolated stereoerrors in the predominantly isotactic poly-
mer to assign in detail the observed signals due to the stereo- and regioirregularities
(Fig. 7.8). This assignment was performed under consideration of the direction of
the polymer chain. Additional GC-analysis of the hydrolyzed polymers has shown a
174 P.T. Altenbuchner et al.

Fig. 7.9 Syndiotactic HH and TT in (R)-PPC synthesized using 1, resulting from “abnormal”
epoxide ring opening at the methine position [78]

correlation between the extent of regioerror and the ring opening of PO with
inversion of configuration at chiral atom, leading to s-HH diads. Signals of
i-diads in the HH region were also observed in certain cases, which is referred to
an “abnormal” PO-enchainment with retention of configuration of the stereocenters.
The assignment of carbonate carbon signals at δC 153.8 and 154.8 ppm to HH and
TT syndiotactic diads, respectively (Fig. 7.9), is consistent with the previous work
of Chisholm and coworkers [80].
However, in the HT-region, a new low-intensity peak in the vicinity of the
ii-signal at the lower field was observed, which could be assigned to an effect of
an HH regioerror close to the isotactic triad. Accordingly, all observed
low-intensity signals in the polymers with a lower isotacticity (δC 154.29, 154.30,
154.35, and 154.39 ppm) correlate well with the resonances in the HH region due to
the regioerror (Fig. 7.10). Such correlation was also observed in case of the polymer
with a predominantly syndiotactic structure, revealing a low-intensity peak at low
field close to an ss-HT-signal.
Aside from the HT region, the splitting pattern of the signals in the HH region of
the spectrum is informative with respect to the copolymer microstructure. The
splitting of the signals in the TT region is not pronounced, therefore indicating a
weak influence of the neighboring stereoconfiguration on the chemical shift of this
junction. It is therefore expected that the HH junction has a stronger effect on
chemical shifts in neighboring stereosequences than the TT sequence does.
In case of copolymers with a high extent of regioerrors, the described correla-
tions are not so straightforward but still can be observed (Fig. 7.11). A greater
number of peaks in the HT and HH regions are expected for such copolymers,
where the model of isolated regioerrors is no longer valid.
Following the (S)-PO/(R)-PO ratios in the polymer chain during the rac-PO/CO2
copolymerization reaction using GC and 500 MHz NMR spectroscopy provided
insight into the understanding of the arising PPC microstructure. The suggested
average microstructure of the copolymers shows a gradient change from iso-
enriched to stereogradient microstructure with the polymerization time due to the
catalyst enantioselectivity. Investigations by Nozaki et al. also led to the synthesis
Fig. 7.10 13C NMR spectra of the carbonate carbon region of (a) iso-enriched PPC (125 MHz,
CDCl3), (b) iso-enriched PPC (225 MHz, CDCl3), and (c) syndio-enriched PPC (225 MHz,
CDCl3) and the proposed assignment of signals on the tetrad level (Reprinted with permission
from Ref. [78]. Copyright 2012, American Chemical Society)

Fig. 7.11 13C (75 MHz, CDCl3) NMR spectra for PPC copolymer synthesized using catalyst
(salen)CrCl with 0.5 equivalent DMAP, (a) using 75 % (S)-PO and 25 % (R)-PO (b) using 25 %
(S)-PO and 75 % (R)-PO at 50  C and 50 bar CO2 for 20 h (Reprinted with permission from Ref.
[78]. Copyright 2012, American Chemical Society)
176 P.T. Altenbuchner et al.

Fig. 7.12 (left) (salen)CoOAc 9 is capable of producing stereogradient PPC, (right) (salen)
CoNO3 10 catalyst with pending TBD arm is capable of producing >99 % HT PPC [76, 81]

of stereogradient PPC which consisted of two enantiomeric structures on each end


of the polymer chain [76]. The used catalyst was an optically active (salen)CoOAc
with pendant ammonium arms. Thermal properties of the polymer were analyzed
by differential scanning calorimetry (DSC) and thermogravimetry (TG).
Different regio- and stereoregularities did not show an influence on the glass
transition temperatures (Tg ~ 33  C) of the obtained polymers as was also con-
firmed by Rieger et al. [78]. But the stereogradient PPC and the stereoblock PPC
exhibited higher thermal decomposition temperatures of up to 273  C compared to
PPCs with lower regio- and stereoregularities and iso-enriched PPC (Td ¼ 240  C).
The authors attributed their findings to a possible stereocomplex formation of a (S)-
PPC block and (R)-PPC block in the same chain upon precipitation in methanol.
Equimolar mixture of (S)-PPC and (R)-PPC did not show increased Td values which
might indicate a spatial dependence of the (S)-PPC and (R)-PPC for the stereoblock
formation. Lu et al. achieved the synthesis of >99 % head-to-tail, >99 % carbonate
linkage, and isotactic PPC with a multichiral (S,S,S)-(salen)CoNO3 catalysts with a
TBD group (1,5,7-triabicylco[4.4.0] dec-5-ene) anchored on the 5-position of one
phenyl ring (Fig. 7.12) [81]. The isotactic PPC produced from (R)-PO exhibited
elevated Tg values of up to 47  C.

7.2.1.5 Microstructure of Poly(cyclohexene carbonate)

The copolymerization of meso-cyclohexene oxide and carbon dioxide produces


poly(cyclohexene oxide) with different stereochemistry depending on the employed
catalyst system. The ring opening of a meso-epoxide proceeds with inversion of the
configuration at one of the two chiral centers (Fig. 7.13).
Nozaki et al. were the first to report asymmetric alternating copolymerization of
meso-CHO with carbon dioxide. They used an equimolar mixture of Et2Zn and (S)-
diphenyl(pyrrolidin-2-yl)methanol as chiral catalyst [82]. Through alki-treatment,
the PCHC was hydrolyzed into trans-1,2-diol and the degree of asymmetric induc-
tion was measured to be 70 % ee. These isotactic PCHCs were characterized via 13C
NMR spectroscopy using statistical methods and model compounds [83]. An
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 177

Fig. 7.13 Poly(cyclohexene oxide): isotactic (a) and syndiotactic (b)

Fig. 7.14 Dinuclear zinc catalyst (11) (left), (salen)CoX catalyst (right) for the copolymerization
of CHO and carbon dioxide [84, 85]

improved dimeric zinc system (11) was able to generate isotactic-enriched PCHC
with 80 % ee [84]. End group analysis showed that the insertion of CO2 happens at
the zinc-ethoxy bond of the catalyst. Nevertheless, the obtained stereoregular
copolymer (80 % ee) showed a Tg of 117  C which is very close to that of
non-stereoregular PCHC (115  C) (Fig. 7.14).
The first syndiotactic-enriched PCHC was reported by Coates et al. who
employed (salen)CoX catalysts for the copolymerization of CHO/CO2 [85]. Cata-
lyst (salen)CoBr (12) and (salen)CoOBzF5 (13) were able to produce PCHC with up
to 81 % r-centered tetrads. Through Bernoullian statistical methods, the PCHC triad
and tetrad sequences were assigned in the 13C NMR spectra in the carbonyl and
methylene regions. The addition of cocatalyst ([PPN]Cl) resulted in the complete
loss in syndiotacticity in the produced PCHC. The synthesis of enantiopure (salen)
Co(III) catalysts (Fig. 7.15) by Lu et al. in 2012 made the synthesis of highly
isotactic PCHC possible [86].
Chiral induction agents, such as (S)-2-methyltetrahydrofuran and (S)-propylene
oxide, further improved the enantioselectivity regarding (S,S)-(salen)Co(III).
PCHC with up to 98:2 RR:SS could be obtained with catalyst 1d/[PPN]Cl in the
presence of (S)-2-methyltetrahydrofuran. The highly isotactic PCHC was analyzed
by DSC measurements and wide angle X-ray diffraction (WAXD). They found that
PCHC samples with less than 90 % isotacticity did not show any crystallization.
178 P.T. Altenbuchner et al.

Fig. 7.15 Disymmetrical enantiopure (salen)Co(III) complexes[86]

(R)-PCHC with an RR:SS ratio of 92:8 ((R)-PCHC-92) gave a Tg of 122  C and a


small endothermic melting peak at 207  C which stems from the low degree of
crystallinity of the polymer. Contrary to amorphous PCHC, the (R)-PCHC-92
sample showed sharp diffraction peaks.
Aside from defined stereo- and regioregular synthesis of optical active PPC and
PCHC, the catalytic side of the polymer formation has received much attention in
recent years. The current development is refocusing on new possible ligand struc-
tures with less toxic metal centers which nevertheless can achieve high activities in
the copolymerization. Those developments will be presented in the next section.

7.3 CO2 Utilization in Polycarbonates

7.3.1 Development

Before diving into the broad field of polycarbonates and their synthesis and applica-
tions, a small review of historic developments is necessary to place the subsequent
work in context. The earliest mention of the synthesis of poly(ethylene carbonate) was
made in a patent in 1966 by Stevens [87]. However, it was not until 1969 that Inoue
et al. presented their findings concerning the activity of a ZnEt2/water mixture which
was able to catalyze the formation of poly(propylene carbonate) from carbon dioxide
and propylene oxide [73]. At pressures ranging from 20 to 50 bar and at 80  C, the
heterogeneous catalyst produced low molecular weight carbonates with turnover
frequencies up to 0.12 h1 and sparked a surge in developments over the next decades.
In heterogeneous catalysis, Soga et al. synthesized the first well-defined catalyst
system from a Zn(OH)2/glutaric acid mixture, which is still the most active zinc-
based system for copolymerization of CO2 and propylene oxide [88]. By using
different zinc carboxylates, Rieger et al. demonstrated the importance of the distance
of the active zinc centers for catalytic activity. Problems arose in the evaluation of
the active zinc centers, as well as the importance of crystallinity, surface, and particle
size for this catalysis [10]. Furthermore, deeper understanding of the reaction mech-
anism was hampered by the drawbacks of heterogeneous catalysts in general. It is very
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 179

challenging to define the active sites in such systems and subsequently to ascribe
structural alterations to activity increases or decreases. Furthermore, the multitude of
active sites in the catalysts leads to broad polydispersity indices (PDI) which makes
the characterization of physical properties of the polymer difficult. Consequently, it
was not until 1978 that Inoue et al. published the first homogeneous catalyst system for
the copolymerization reaction of CO2 and epoxides [89]. The aluminum tetraphenyl-
porphyrin complexes in combination with EtPh3PBr as cocatalyst were able to
copolymerize not only cyclohexene oxide (CHO) but also propylene oxide
(PO) with low PDIs. In 1995, bis(phenoxide)Zn(THF)2 catalysts were synthesized
and employed in the copolymerization of PO and also CHO and CO2 [47, 90]. This
catalyst system exhibited rather low activities and selectivities toward polycarbonate
formation. The next important step toward well-defined and tunable catalysts was
made by Coates et al. with the development of zinc β-diiminate complexes
((β-diiminate)ZnOR) [91]. This type of ligand allowed adjustment of the steric and
electronic environment around the metal center and even small variations in the ligand
framework already had drastic influences on activity and selectivity. Recently, Rieger
et al. published a new dinuclear zinc catalyst system for the copolymerization of
cyclohexene oxide and carbon dioxide. The catalyst 35 limits – in a certain broad
pressure regime – the polymerization rate as a function of the applied CO2 pressure.
It is the first time that the rate determining step is shifted toward CO2 insertion.
This unique behavior is attributed to a flexible CH2-tether that links both complex
moieties. As a result, the highest known polymerization activities for CHO-based
polycarbonates are reported.
Nowadays, development of these well-defined homogeneous catalysts and
research on polycarbonates from carbon dioxide and epoxides are flourishing.
Various review articles have been published in recent years in this field of research
which demonstrates the growing interest in this material and the underlying mech-
anism [5, 9, 88, 92–97].

7.3.2 Mechanistic Aspects of the Epoxide and Carbon


Dioxide Copolymerization

Homogeneous catalysts with the general structure LnMX possess the great advan-
tage of having only one defined active site. Tailoring the organic ligand Ln as well
as the initiating group and the metal center provides a broad range of possibilities.
Nevertheless, mechanistic investigations so far have not provided an unambiguous
picture of the copolymerization of carbon dioxide and epoxides. Hence, three
different possible reaction pathways shall be presented for a model salen complex.
The reaction pathway A requires interactions between two catalyst molecules and
their active sites (Fig. 7.16). Thus, this mechanism is probable in the absence of a
cocatalyst at low epoxide to catalyst loadings. It is rate dependent on the Lewis acidity
of the corresponding metal center and on the nucleophilicity of the axial ligand X.
180 P.T. Altenbuchner et al.

Fig. 7.16 Initiation mechanism for LnMX for the copolymerization of carbon dioxide and PO:
bimetallic pathway (a), monometallic pathway (b), and binary pathway (c) (Reprinted with
permission from Ref. [88]. Copyright 2011, Elsevier)

Accordingly, the type of catalytic system used defines the mechanism. In the work of
Jacobsen on the asymmetric nucleophilic ring opening of epoxides by chiral (salen)CrX
complexes, this intermolecular, bimetallic pathway is a vital step in the absence of a
cocatalyst [98]. The group of Rieger et al. also performed theoretical calculations
concerning the chain-growth mechanism during the copolymerization of CO2 and
epoxides. The DFT calculations indicated that chain growth proceeds through the attack
of a metal-bound alkyl carbonate on a precoordinated epoxide on a metal center
[99]. However, other investigations indicated a bimetallic initiation step and subsequent
monometallic chain propagation [71, 100–102]. Reaction pathway B is the monome-
tallic mechanism and is comparable to an associative ligand exchange mechanism. In
this scenario, the nucleophile and the epoxide are both bound to one metal center.
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 181

Fig. 7.17 Bimetallic propagation and the possible concomitant side reactions

In general, the epoxide is opened at the least hindered C-O bond but regioerrors may
occur. Both the axial ligand and the nucleophile have the potential to open the epoxide
and are therefore incorporated in the polymer chain as end groups. The thermally
disfavored transition state makes this reaction mechanism unlikely for commonly
used catalysts. Reaction pathway C takes place in the case of binary catalyst/cocatalyst
systems where the added nucleophile attacks the precoordinated epoxide and opens the
ring. Binary catalyst/cocatalyst systems have been under scrutiny from multiple
research teams around the world in recent times [46, 85, 103–116].
The group of Coates et al. was able to show with (β-diiminate)ZnOR catalysts
that a bimetallic mechanism is at play during their reactions [117]. Furthermore,
they were able to find the rate determining step as the incorporation of the
epoxide. The weak nucleophilic nature of the carbonate end group makes a
preactivation of the epoxide necessary. Therefore, Lewis acids in the form of
cocatalysts or a second metal center have to be used. After the ring opening of the
epoxide, CO2 is inserted into the metal alkoxy bond, for which an open coordi-
nation site at the metal is not obligatory. This mechanism explains the generally
observed loss of activity of cocatalyst systems at higher dilutions (Fig. 7.17).
Undesired side reactions during the copolymerization are the formation of cyclic
carbonates and the formation of polyether segments. The group of Darensbourg
182 P.T. Altenbuchner et al.

Fig. 7.18 Chain backbiting mechanism [88]

studied in detail the reasons for the product selectivity and the dependence on
temperature, pressure, cocatalyst, and catalyst structure [101]. In general, both low
temperatures and stronger coordination to the metal center decrease the amount of
cyclic byproduct since these factors can suppress backbiting. During the backbiting
reaction either the carbonato- or alkoxy-chain end of the polymer attacks at the
growing chain (Fig. 7.18). This reaction is especially pronounced for aliphatic
epoxides [63]. Alternatively, the growing chain end may dissociate from the
metal center and more easily undergo backbiting in its unbound state [99]. The
protonation of this unbound alkoxy- or carbonato-chain end is a possible strategy to
reduce backbiting.
Chain transfer reactions during the copolymerization can be caused by traces of
water, alcohols, or acids. Chain transfer can be exploited to tune the molecular weight
by addition of, for example, adipic acid or telechelic polymers featuring alcohol end
groups [118]. These chain transfer reactions lead to lower molecular weights than
theoretically calculated. Furthermore, phosphoric acid and phenylphosphonic acids
as additives have been tested to afford flame-retarding PPCs [42].

7.3.3 Catalysis

7.3.3.1 Bifunctional Catalysts

Owing to the desire to reach high selectivities and activities, catalyst design has
shifted in recent years toward bifunctional catalysts. These complexes combine the
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 183

Fig. 7.19 Copolymerization mechanism for a bifunctional salenCo(III) complex with a


piperidinium arm [67] (Reprinted with permission from Ref. [88]. Copyright 2011, Elsevier)

active metal site, coordinated by an organic ligand framework, with a covalently


bound cocatalyst. With these considerations in mind Nozaki et al. incorporated a
piperidinium arm into a (salen)CoOAc complex. It was proposed that cyclic
carbonate formation is inhibited by the protonated piperidinium arm which can
protonate the growing polymer chain upon dissociation from the metal center
(Fig. 7.19). At a catalyst loading of PO/catalyst ¼ 2,000, 14 bar and ambient
temperatures, a selectivity of 99 % for poly(propylene carbonate) was achieved.
Furthermore, it was shown that even at elevated temperatures of 60  C, the cyclic
carbonate formation could be suppressed to a certain degree [67].
The group of Lee et al. designed a series of ionic bifunctional catalyst structures
in which the cocatalyst is covalently bound and is supposed to keep the growing
polymer chain close to the reactive center [22, 119]. The anions also function as
initiators for the copolymerization and have a pronounced effect on polymerization
activity and product selectivity [120]. These catalysts were employed in copoly-
merization reactions with various epoxides, for example, cyclohexene oxide,
hexene oxide, 1-butene oxide, and propylene oxide [48]. Polymerizations of PO
and carbon dioxide at 80  C and at dilutions as low as 0.67 mmol% were shown to
be possible while retaining high activities and selectivities for the polycarbonate
formation (TOF 12,400 h1, 96 %) [22]. The highest reported TOF is 26,000 h1
(>99 %, Mn ¼ 114 kg/mol, PDI ¼ 1.29). It was possible to produce high molec-
ular weight PPC at elevated temperatures with a Mn of up to 285 kg/mol and narrow
molecular weight distributions of 1.18. Interestingly, the catalyst structure allowed
the removal from the produced polymer after the copolymerization.
184 P.T. Altenbuchner et al.

Fig. 7.20 Bifunctional ionic salenCo(III) catalyst 5 and its proposed bidentate coordination
structure [121]

Filtration through a short pad of silica gel results in a catalyst which exhibits
almost equivalent copolymerization activities even after several separation cycles.
In their later work the group of Lee discovered an unusual coordination mode in
their bifunctional catalysts. Though no crystal structure has yet been obtained from
this proposed conformation, NMR studies and DFT calculations support the pro-
posed structure. They postulated that the imine nitrogens of the salen ligand do
not coordinate to the cobalt center (Fig. 7.20). Instead the DNPs coordinate to
the cobalt, forming a negatively charged cobaltate complex. Structural variations of
the ligand framework allowed the conclusion that sterically more demanding
substituents in ortho-position reduce the activity. The authors attributed this to
the fact that the formation of the aforementioned bidentate cobalt coordination is
inhibited. High copolymerization activities were only observed on catalysts with
the ability to form the cobaltate complex (Fig. 7.20 (right)) [121]. The employed
2,4-dinitrophenolate is explosive in its dry state as well as hazardous. Therefore,
anion variations were performed using 2,4,5-trichlorophenolate, 4-nitrophenolate,
2,4-dichlorophenolate, and nitrate [118, 120]. It was found by chance that
homoconjugated anion pairs of the phenols [X-H-Y] show increased activities
and higher tolerance to water impurities leading to shorter induction periods
[121]. Interestingly the rather long induction periods of the cobalt complexes during
the copolymerization of epoxides and carbon dioxide are still unexplained. Further
examples of bifunctional catalysts for the production of polycarbonates were
published by Lu et al. These (salen)CoX complexes feature either an alkyl arm
with pending quaternary ammonium salt at the 3-position of the phenyl ring or an
alkyl arm bound to 1,5,7-triabicyclo[4,4,0] dec-5-ene (TBD). Catalyst 4 with the
TBD group was able to polymerize with high selectivity for polycarbonate (97 %)
at 100  C and 20 bar. Activities up to 10,880 h1 at a loading of 1 mmol% were
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 185

Fig. 7.21 Bifunctional salenCo(III) catalysts [22, 66, 67]

reached [122]. Through mass spectroscopy, it was possible to show that the pending
TBD arm is activated after the insertion of one PO and carbon dioxide molecule. It
stabilizes the Co(III) species against reduction to the inactive Co(II) by reversible
intramolecular Co-O bond formation (Fig. 7.21).
The Co(III) catalyst (3) with an ammonium salt (-NEt2Me+) and
2,4-dinitrophenolate proved to be less active with TOF of 3,900 h1 at 90  C
while nevertheless maintaining high selectivity for copolymer formation.
Bifunctional cobalt catalysts are currently the catalysts of choice for further
scientific studies and improvements due to their excellent activities combined with
high selectivities for the polymer formation. However, their main drawbacks cannot
be solved by synthetic design as cobalt is toxic, prone to reduction, and has to be
186 P.T. Altenbuchner et al.

removed from the polymer after copolymerization. This puts a considerable ener-
getic burden on the produced polymer and hampers its chances to compete with
already established polymers.

7.3.3.2 Zinc Catalysts

Due to their colorlessness, nontoxicity, and low price, it is not surprising that there
is a considerable amount of literature on Zn(II)-based catalysts. For heterogeneous
catalysis, the zinc glutarate (ZnGA) system is the most widely studied one. It is easy
to prepare, nontoxic, economically viable, and easy to handle. But these heteroge-
neous catalysts are not only interesting for industrial applications as they offer the
opportunity to better understand the reaction processes. The molecular structure of
ZnGA was only recently reported and showed a unique structural constitution
[123]. It consists of Zn centers coordinated by four carboxyl groups with the
glutarate ligands either in a bent or an extended conformation. This structural
type limits the activity for the copolymerization to the surface due to restricted
monomer diffusion [124]. Consequently, methods for enlarging the surface were
tested but with unsatisfactory results with respect to the achieved activity. Also the
replacement of GA by derivatives (e.g., 2-ketoglutaric acid, 3,3-dimethylglutaric
acid) with other functionalities in the hydrocarbon chain did not bring the expected
success [125]. Experiments with different alkyl lengths (succinic acid (SA), adipic
acid (AA), pimelic acid (PA)) in between the Zn centers showed that these are also
active [126, 127]. It was not until 2011 that Rieger et al. published a detailed study
of ZnSA, ZnGA, ZnAA, and ZnPA and their respective activities in the copoly-
merization of epoxides and CO2 [128]. From the copolymerization experiments, the
activity gap between ZnSA and the higher homologues was identified. The molec-
ular structure was the crucial factor as the main difference between ZnSA and the
tested homologues was the Zn-Zn distance. This indicated that the defined spatial
structure which is influenced by the dicarboxylic acid has direct influence on the
activity of heterogeneous zinc dicarboxylate systems. The Zn-Zn distance that is
necessary for copolymerization is only found on one of the main hkl-indiced plains
of the solid-state structure of ZnSA. However, in the solid-state structure of ZnGA,
the corresponding distance of 4.6–4.8 Å is found on each main hkl plain (Fig. 7.22).
These results show that two zinc centers have to be in close spatial proximity for
high copolymerization activities to be possible. Theoretical calculations suggest
that the optimal Zn-Zn distance for CO2/epoxide copolymerization is between 4.3
and 5.0 Å [129].
This range gives an optimum between the activation energy and product selec-
tivity toward polymer formation. The importance of the appropriate Zn-Zn distance
for homogeneous catalysts was already described in literature. The study of Rieger
et al. indicates that also in heterogeneous Zn catalysts a bimetallic mechanism is at
play on the surface for which the spatial separation of the metal is crucial. To gain
deeper insights into the copolymerization mechanism, homogeneous single-site
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 187

Fig. 7.22 Close-up view on unit cell and metal distances in zinc glutarate (Zn-Zn distance ¼ 4.6-
4.8 Å) (Reprinted with permission from Ref. [128]. Copyright 2011, American Chemical Society)

catalysts are required. Darensbourg et al. synthesized the first homogeneous zinc-
based catalyst for the copolymerization of cyclohexene oxide and CO2 [47]. These
catalysts exhibited a strong tendency to form polyether even at elevated CO2
pressures (55 bar). Major drawbacks of those early systems were the low activities
and selectivities for polycarbonate formation. In the case of the phenoxide zinc
complexes, another side reaction limited success as their phenoxide ligands were
consumed as initiators during the polymerization [51, 130]. The unstable nature of
the zinc phenoxide catalysts made mechanistic investigations and the defined
synthesis of the ligand framework around the metal center difficult. The first
breakthrough for copolymerization of epoxides and carbon dioxide was the discov-
ery of β-diketiminato zinc catalysts (BDI) for the CHO/CO2 copolymerization by
Coates et al. which led to a more systematic catalyst design. Minor variations in the
electronic and steric character of the BDI ligand framework resulted in dramatic
changes in catalytic activity [47, 48]. These catalysts afforded polymers with
narrow PDIs at low carbon dioxide pressures with high activities. In the copoly-
merization reaction of CHO/CO2, minor changes in the backbone from R1 ¼ H, R2,
R3 ¼ Et to R1 ¼ CN, R2 ¼ Me and R3 ¼ iPr led to an activity increase from
239 to 2,290 h1 [91]. As shown in Fig. 7.23, these complexes are present in
monomer/dimer equilibria which depend on sterics, electronics, and temperature.
188 P.T. Altenbuchner et al.

Fig. 7.23 Monomer/dimer equilibria for the zinc-based catalysts by Coates et al. [91]

Rate studies with in situ IR spectroscopy on the copolymerization resulted in a


zero-order dependence in CO2, a first-order dependence in CHO, and a varying
dependency from 1.0 to 1.8 for zinc [117]. This variation can be constituted with the
appearance of the mentioned equilibrium in Fig. 7.23. All results support the
assumption that two zinc centers are interacting in the copolymerization and that
the rate determining step is located at the ring opening of the epoxide. Additionally,
it was found that sterical variation (R1 ¼ Et, R2 ¼ iPr) at the aniline moiety and the
introduction of electron withdrawing groups to the catalyst (20) backbone gave an
active catalyst for the copolymerization of PO/CO2 at 25  C, 6.9 bar with a TOF of
235 h1. The resulting polymer featured a regioirregularity and a narrow molecular
weight distribution [131] (Fig. 7.24).
Interestingly this complex structure also polymerized β-butyrolactone and
β-valerolactone to afford atactic poly(3-hydroxybutyrate) (PHB) and poly
(3-hydroxyvalerate) [132]. Research into the zinc BDI systems led to the discovery
of ethylsulfinate as superior initiating group [133]. In 2005 Lee et al. presented an
anilido-aldimine ligand for dinuclear zinc complexes (Fig. 7.25). Thereby, open
(22) and closed (21–24) structures were synthesized with zinc-zinc distances of
between 4.88 Å for the open structure and 4.69 Å for the closed structure. For the
copolymerization of cyclohexene oxide and carbon dioxide, only the catalyst with
the open structure was active. With this catalyst TOFs up to 200 h1 were achieved.
Increasing the Lewis acidity of the zinc centers led to an increase of activity up to
2,860 h1 at high dilutions ([catalyst]:[CHO] ¼ 1:50,000) [134, 135].
In 2011, Williams et al. published a rigid dinuclear zinc complex for low-pressure
copolymerization of cyclohexene oxide and CO2. Investigations of this catalyst
system also revealed a zero-order dependence on CO2 (between 1 and 40 bar CO2
pressure) and a first-order dependence on CHO. These results confirmed the hypoth-
esis that the incorporation of the epoxide is rate determining for their dinuclear rigid
zinc catalysts [136]. This implies that at typical process conditions, an increase in
CO2 concentration does not automatically enhance polymerization activity. In sub-
sequent work, Williams et al. replaced zinc with cobalt(II)/(III) [137] and iron [138]
in the reduced Robson-type ligand structure (Fig. 7.26). The isolated complexes
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 189

Fig. 7.24 Zinc catalyst for copolymerization of propylene oxide and carbon dioxide [131]

Fig. 7.25 Dinuclear anilido-aldimine zinc complexes [134, 135]

were all able to copolymerize CHO and carbon dioxide with good activities and high
selectivities. The highest activities in the copolymerization of CHO and CO2 were
achieved with magnesium (31). Activities of up to 750 h1 and very high selectivities
of >99 % for the copolymer formation were observed with this structure (12 atm
CO2, [CHO]:[31] ¼ 1:10,000, PDI ¼ 1.03/1.10). The data gathered from these metal
screenings show very nicely how the copolymerization activity can be tuned in
dinuclear catalysts by the Lewis acidity of the metal itself.
In 2013, Rieger et al. presented a flexibly tethered dinuclear zinc complex 35, in
which two BDI zinc units are linked by a tether (Fig. 7.27). The linker allows to
overcome the entropically disfavored aggregation of two individual complex mol-
ecules in solution. Catalyst 35 exhibits very high activities of up to 9,130 h1 for the
190 P.T. Altenbuchner et al.

Fig. 7.26 Dinuclear reduced Robson-type complexes with Zn-, Mg-, Co-, and Fe-centers
[64, 136–139]

Fig. 7.27 Flexibly tethered


dinuclear zinc complex
[140]

copolymerization of CHO and carbon dioxide (at 40 bar CO2, 100  C, [CHO]:
[35] ¼ 4,000:1) [140].
Kinetic measurements in an in situ IR reactor revealed a first-order dependence
in catalyst. Interestingly, with constant catalyst and cyclohexene oxide concentra-
tion, the order in carbon dioxide was determined to be one in the range of 5–25 bar
CO2 pressure. For 25–45 bar, this order changes from one to zero. The assignment
of the order in cyclohexene oxide was performed at two different pressure regimes:
10 and 30 bar, respectively. This resulted in a reaction order of zero for 10 bar and a
reaction order of one at 30 bar. The subsequent rate laws are depicted below:
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 191

Fig. 7.28 Enthalpy and Gibbs free energy profile for the whole catalytic cycle of catalyst 35
(Reprinted with permission from Ref. [140]. Copyright 2013, Wiley-VCH Verlag GmbH &
Co. KGaA)

Equation 7.1: Rate law of complex 35 for 5–25 bar CO2 (a) and for 25–45 bar
CO2 (b) [140].

r ¼ k  ½CHO0  ½CO2 1  ½Catalyst1 525 bar CO2 ðaÞ


1
r ¼ k  ½CHO  ½CO2   ½Catalyst
0 1
2545 bar CO2 ð bÞ

Catalyst 35 is the first dinuclear zinc catalyst which shows a shift in the rate
determining step from ring opening of the epoxide to carbon dioxide insertion for
the copolymerization of cyclohexene oxide and carbon dioxide in a certain pressure
regime. This unique behavior is attributed to the flexible CH2-tethers which links
both complex moieties. Considering the first-order dependence with respect to
CHO of all other catalysts, this suggests that the flexible tether between the two
zinc centers facilitates the cooperative ring-opening step in an unprecedented
manner. These results are supported by quantum chemical calculations. The com-
puted results for the main steps, beginning from the dicarbonato complex as the
starting point, followed by coordination of CHO, epoxide ring opening to form an
alkoxide and subsequent insertion of CO2 into the coordinated alkoxide bond, are
summarized in Fig. 7.28.
192 P.T. Altenbuchner et al.

As can be seen in Fig. 7.28, both elementary steps give similar computed G-values
(~100 kJ/mol) which means that both transition states are equally difficult to over-
come. It is notable that the enthalpy H is much bigger for the ring opening (TS
(a) ¼ 52.3 kJ/mol) compared to the H of the CO2 insertion (TS(b) ¼ 19.2 kJ/mol).
At first glance, it appears surprising that elementary steps with such dissimilar
activation energies can both be rate limiting. DFT calculations explain the
similar activation energies as a consequence of Gibbs free energy which includes
both activation enthalpy and entropy. The main point is that all alkoxide species
are significantly higher in G if compared to the dicarbonato complex I, which is
why all Gibbs free energies have to be considered in relation to complex I. As the
resting state is in all cases the dicarbonato complex I, oxirane ring opening is a
bimolecular reaction, i.e., the incorporation of one liquid species into the catalyst
bound polymer chain. Otherwise, CO2 insertion is effectively a trimolecular
reaction, which consumes both one liquid epoxide and one gaseous CO2. All the
structural and electronic features of 35 lead to high copolymerization activities of
over 9,130 h1 and selectivities of >99 % toward polycarbonate formation.
In conclusion, zinc and other nontoxic, cheap, and abundant metals have tre-
mendous advantages compared to other transition metals, namely, cobalt and
chrome, which are currently predominantly employed for the copolymerization of
carbon dioxide and epoxides. In recent publications, it was nicely shown how those
economic and eco-friendly elements can achieve high activities and selectivities.
An additional advantage of zinc and magnesium is their colorless ions. This opens
up the possibility to leave the catalyst in the formed polymer and reduce work-up
costs.

7.4 Conclusion

Carbon dioxide as feedstock for the generation of biodegradable polycarbonates has


been known in the literature for almost 45 years. Since the discovery of the
copolymerization reaction of carbon dioxide and epoxides, chemistry has made
significant progress, and especially in the last 10 years, the search for better and
more eco-friendly catalysts has gained momentum. As the stereoselective synthesis
of poly(propylene carbonate) is still in its infancy and the analytical methods used
have not yet been fully exploited, NMR techniques have the potential to help
interpret the structures produced by catalyst systems and to pinpoint promising
catalyst candidates for further development. Though heterogeneous catalysts still
dominate industrial production processes, for mechanistic investigations, homoge-
neous systems do have the necessary advantages to tackle the underlying questions
about the mechanism and within the catalysis. Cobalt(III) catalysts with a variety of
accompanying ligand structures have become the center of attention since the
discovery of bifunctional structures which allow the incorporation of the cocatalyst
into the complex and thereby enable high activities and the production of
polycarbonates with high molecular masses, low PDIs, and >99 % carbonate
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 193

content. However, cobalt as toxic and coloring metal has to be removed from the
polymer after copolymerization. Any additional work-up increases the energetic
burden on production and limits competitiveness of biodegradable polymers such
as poly(propylene carbonate). Possible applications in food packaging also make
the development of less toxic catalysts desirable. An applicable catalyst for large-
scale industrial production in a continuous process has to exhibit a sufficient
activity to make it economically viable. In recent years, research made great
progress in the development of iron-, magnesium-, and zinc-based systems which
show improved activities. It will be interesting and exciting to see the future
development of catalyst systems which tolerate water contaminations and multiple
epoxides and nevertheless show high activities in co- and terpolymerizations at
ambient conditions.

References

1. Song C (2006) Global challenges and strategies for control, conversion and utilization of CO2
for sustainable development involving energy, catalysis, adsorption and chemical processing.
Catal Today 115:2–32
2. Energy, E.A.U.S.D.o. (2011) Annual Energy Review. http://www.eia.gov/totalenergy/data/
annual/index.cfm. Accessed 02 May 2013
3. Tans P (2013) Trends in atmospheric carbon dioxide. www.esrl.noaa.gov/gmd/ccgg/trends/.
Accessed 02 May 2013
4. Sakakura T, Kohno K (2009) The synthesis of organic carbonates from carbon dioxide. Chem
Commun (Cambridge, UK) 11:1312–1330
5. Cokoja M, Bruckmeier C, Rieger B et al (2011) Transformation of carbon dioxide with
homogeneous transition-metal catalysts: a molecular solution to a global challenge? Angew
Chem Int Ed 50(37):8510–8537
6. Sakakura T, Choi JC, Yasuda H (2007) Transformation of carbon dioxide. Chem Rev
107:2365–2387
7. Peters M, Koehler B, Kuckshinrichs W et al (2011) Chemical technologies for exploiting and
recycling carbon dioxide into the value chain. ChemSusChem 4:1216–1240
8. Darensbourg DJ, Wilson SJ (2012) What’s new with CO2? Recent advances in its copoly-
merization with oxiranes. Green Chem 14(10):2665–2671
9. Lu X-B, Darensbourg DJ (2012) Cobalt catalysts for the coupling of CO2 and epoxides to
provide polycarbonates and cyclic carbonates. Chem Soc Rev 41:1462–1484
10. Clements JH (2003) Reactive applications of cyclic alkylene carbonates. Ind Eng Chem Res
42:663–674
11. North M, Pasquale R, Young C (2010) Synthesis of cyclic carbonates from epoxides and
CO2. Green Chem 12(9):1514–1539
12. Varghese JK, Na SJ, Park JH et al (2010) Thermal and weathering degradation of poly
(propylene carbonate). Polym Degrad Stab 95(6):1039–1044
13. Du LC, Meng YZ, Wang SJ et al (2004) Synthesis and degradation behavior of poly
(propylene carbonate) derived from carbon dioxide and propylene oxide. J Appl Polym Sci
92:1840–1846
14. Luinstra GA (2008) Poly(propylene carbonate), old copolymers of propylene oxide and
carbon dioxide with new interests: catalysis and material properties. Polym Rev (Philadel-
phia, PA, USA) 48(1):192–219
15. Qin Y, Wang X (2010) Carbon dioxide-based copolymers: environmental benefits of PPC, an
industrially viable catalyst. Biotechnol J 5:1164–1180
194 P.T. Altenbuchner et al.

16. Barreto C, Hansen E, Fredriksen S (2012) Novel solventless purification of poly(propylene


carbonate): tailoring the composition and thermal properties of PPC. Polym Degrad Stab
97(6):893–904
17. Chisholm MH, Navarro-Llobet D, Zhou Z (2002) Poly(propylene carbonate). 1. More about
poly(propylene carbonate) formed from the copolymerization of propylene oxide and carbon
dioxide employing a zinc glutarate catalyst. Macromolecules 35:6494–6504
18. Kuran W, Gorecki P (1983) Degradation and depolymerization of poly(propylene carbonate)
by diethylzinc. Makromol Chem 184:907–912
19. Gorecki P, Kuran W (1985) Diethylzinc-trihydric phenol catalysts for copolymerization of
carbon dioxide and propylene oxide: activity in copolymerization and copolymer destruction
processes. J Polym Sci Polym Lett Ed 23:299–304
20. Darensbourg DJ, Wei S-H, Wilson SJ (2013) Depolymerization of poly(indene carbonate).
A unique degradation pathway. Macromolecules 46:3228–3233
21. Wang SJ, Du LC, Zhao XS et al (2002) Synthesis and characterization of alternating
copolymer from carbon dioxide and propylene oxide. J Appl Polym Sci 85(11):2327–2334
22. Min SSJK, Seong JE, Na SJ et al (2008) A highly active and recyclable catalytic system for
CO2/propylene oxide copolymerization. Angew Chem Int Ed 47(38):7306–7309
23. Peng S, An Y, Chen C et al (2003) Thermal degradation kinetics of uncapped and end-capped
poly(propylene carbonate). Polym Degrad Stab 80:141–147
24. Dixon DD, Ford ME, Mantell GJ (1980) Thermal stabilization of poly(alkylene carbonate)s.
J Polym Sci Polym Lett Ed 18(2):131–134
25. Lai MF, Li J, Liu JJ (2005) Thermal and dynamic mechanical properties of poly(propylene
carbonate). J Therm Anal Calorim 82:293–298
26. Yu T, Luo F-L, Zhao Y et al (2011) Improving the processability of biodegradable polymer
by stearate additive. J Appl Polym Sci 120:692–700
27. Yao M, Mai F, Deng H et al (2011) Improved thermal stability and mechanical properties of
poly(propylene carbonate) by reactive blending with maleic anhydride. J Appl Polym Sci
120:3565–3573
28. Yu T, Zhou Y, Zhao Y et al (2008) Hydrogen-bonded thermostable liquid crystalline complex
formed by biodegradable polymer and amphiphilic molecules. Macromolecules
41(9):3175–3180
29. Ge XC, Zhu Q, Meng YZ (2006) Fabrication and characterization of biodegradable poly
(propylene carbonate)/wood flour composites. J Appl Polym Sci 99:782–787
30. Bian J, Wei XW, Lin HL et al (2011) Preparation and characterization of modified graphite
oxide/poly(propylene carbonate) composites by solution intercalation. Polym Degrad Stab
96:1833–1840
31. Shi X, Gan Z (2007) Preparation and characterization of poly(propylene carbonate)/mont-
morillonite nanocomposites by solution intercalation. Eur Polym J 43:4852–4858
32. Wang JT, Zhu Q, Lu XL et al (2005) ZnGA–MMT catalyzed the copolymerization of carbon
dioxide with propylene oxide. Eur Polym J 41(5):1108–1114
33. Zhang Z, Lee J-H, Lee S-H et al (2008) Morphology, thermal stability and rheology of poly
(propylene carbonate)/organoclay nanocomposites with different pillaring agents. Polymer
49:2947–2956
34. Chen L, Qin Y, Wang X et al (2011) Toughening of poly(propylene carbonate) by
hyperbranched poly(ester-amide) via hydrogen bonding interaction. Polym Int 60:1697–1704
35. Ma X, Yu J, Wang N (2005) Compatibility characterization of poly(lactic acid)/poly(propyl-
ene carbonate) blends. J Polym Sci B 44:94–101
36. Wang N, Zhang X, Yu J et al (2008) Partially miscible poly(lactic acid)-blend-poly(propyl-
ene carbonate) filled with carbon black as conductive polymer composite. Polym Int
57:1027–1035
37. Peng S, Wang X, Dong L (2005) Special interaction between poly(propylene carbonate) and
corn starch. Polym Compos 26:37–41
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 195

38. Spencer TJ, Chen Y-C, Saha R et al (2011) Stabilization of the thermal decomposition of poly
(propylene carbonate) through copper ion incorporation and use in self-patterning. J Electr
Mater 40:1350–1363
39. Yu T, Zhou Y, Liu K et al (2009) Improving thermal stability of biodegradable aliphatic
polycarbonate by metal ion coordination. Polym Degrad Stab 94:253–258
40. Uzunlar E, Kohl PA (2012) Thermal and photocatalytic stability enhancement mechanism of
poly(propylene carbonate) due to Cu(I) impurities. Polym Degrad Stab 97:1829–1837
41. Jeon JY, Lee JJ, Varghese JK et al (2012) CO2/ethylene oxide copolymerization and ligand
variation for a highly active salen-cobalt(iii) complex tethering 4 quaternary ammonium
salts. Dalton Trans 42:9245–9254
42. Cyriac A, Lee SH, Varghese JK et al (2011) Preparation of flame-retarding poly(propylene
carbonate). Green Chem 13(12):3469–3475
43. Cherian AE, Sun FC, Sheiko SS et al (2007) Formation of nanoparticles by intramolecular
cross-linking: following the reaction progress of single polymer chains by atomic force
microscopy. J Am Chem Soc 129:11350–11351
44. Cyriac A, Lee SH, Lee BY (2011) Connection of polymer chains using diepoxide in CO2/
propylene oxide copolymerizations. Polym Chem 2(4):950–956
45. Darensbourg DJ, Poland RR, Strickland AL (2012) (Salan)CrCl, an effective catalyst for the
copolymerization and terpolymerization of epoxides and carbon dioxide. J Polym Sci Part A
Polym Chem 50(1):127–133
46. Shi L, Lu X-B, Zhang R et al (2006) Asymmetric alternating copolymerization and
terpolymerization of epoxides with carbon dioxide at mild conditions. Macromolecules
39:5679–5685
47. Darensbourg DJ, Holtcamp MW (1995) Catalytic activity of zinc(II) phenoxides which
possess readily accessible coordination sites. Copolymerization and terpolymerization of
epoxides and carbon dioxide. Macromolecules 28(22):7577–7579
48. Seong JE, Na SJ, Cyriac A et al (2009) Terpolymerizations of CO2, propylene oxide, and
various epoxides using a cobalt(III) complex of salen-type ligand tethered by four quaternary
ammonium salts. Macromolecules 43(2):903–908
49. Harrold ND, Li Y, Chisholm MH (2013) Studies of ring-opening reactions of styrene oxide
by chromium tetraphenylporphyrin initiators. Mechanistic and stereochemical consider-
ations. Macromolecules 46(3):692–698
50. Kim JG, Cowman CD, LaPointe AM et al (2011) Tailored living block copolymerization:
multiblock poly(cyclohexene carbonate)s with sequence control. Macromolecules
44(5):1110–1113
51. Darensbourg DJ, Wildeson JR, Yarbrough JC et al (2000) Bis 2,6-difluorophenoxide dimeric
complexes of zinc and cadmium and their phosphine adducts: lessons learned relative to
carbon dioxide/cyclohexene oxide alternating copolymerization processes catalyzed by zinc
phenoxides. J Am Chem Soc 122(50):12487–12496
52. Wu G-P, Wei S-H, Lu X-B et al (2010) Highly selective synthesis of CO2 copolymer from
styrene oxide. Macromolecules 43(21):9202–9204
53. Wu G-P, Xu P-X, Lu X-B et al (2013) Crystalline CO2 copolymer from epichlorohydrin via
Co(III)-complex-mediated stereospecific polymerization. Macromolecules 46(6):2128–2133
54. Darensbourg DJ, Wilson SJ (2011) Synthesis of poly(indene carbonate) from indene oxide
and carbon dioxide – a polycarbonate with a rigid backbone. J Am Chem Soc
133(46):18610–18613
55. Liu S, Xiao H, Huang K et al (2006) Terpolymerization of carbon dioxide with propylene
oxide and ε-caprolactone: synthesis, characterization and biodegradability. Polym Bull
56:53–62
56. Liu Y, Huang K, Peng D et al (2006) Synthesis, characterization and hydrolysis of an
aliphatic polycarbonate by terpolymerization of carbon dioxide, propylene oxide and maleic
anhydride. Polymer 47(26):8453–8461
196 P.T. Altenbuchner et al.

57. Hwang Y, Jung J, Ree M et al (2003) Terpolymerization of CO2 with propylene oxide and
ε-caprolactone using zinc glutarate catalyst. Macromolecules 36(22):8210–8212
58. Lu L, Huang K (2005) Synthesis and characteristics of a novel aliphatic polycarbonate, poly
[(propylene oxide)-co-(carbon dioxide)-co-(γ-butyrolactone)]. Polym Int 54:870–874
59. Liu S, Wang J, Huang K et al (2011) Synthesis of poly(propylene-co-lactide carbonate) and
hydrolysis of the terpolymer. Polym Bull 66:327–340
60. Wu G-P, Darensbourg DJ, Lu X-B (2012) Tandem metal-coordination copolymerization and
organocatalytic ring-opening polymerization via water to synthesize diblock copolymers of
styrene oxide/CO2 and lactide. J Am Chem Soc 134:17739–17745
61. Zhang J-F, Ren W-M, Sun X-K et al (2011) Fully degradable and well-defined brush
copolymers from combination of living CO2/epoxide copolymerization, thiol-ene click
reaction and ROP of ε-caprolactone. Macromolecules 44:9882–9886
62. Kröger M, Folli C, Walter O et al (2005) Alternating copolymerization of cyclohexene oxide
and CO2 catalyzed by zinc complexes with new 3-amino-2-cyanoimidoacrylate ligands. Adv
Synth Catal 347(10):1325–1328
63. Lehenmeier MW, Bruckmeier C, Klaus S et al (2011) Differences in reactivity of epoxides in
the copolymerisation with carbon dioxide by zinc-based catalysts: propylene oxide versus
cyclohexene oxide. Chem Eur J 17(32):8858–8869
64. Kember MR, Knight PD, Reung PTR et al (2009) Highly active dizinc catalyst for the
copolymerization of carbon dioxide and cyclohexene oxide at one atmosphere pressure.
Angew Chem Int Ed 48:931–933
65. Seong JE, Na SJ, Cyriac A et al (2010) Terpolymerizations of CO2, Propylene oxide, and
various epoxides using a cobalt(III) complex of salen-type ligand tethered by four quaternary
ammonium salts. Macromolecules 43(2):903–908
66. Ren W-M, Zhang X, Liu Y et al (2010) Highly active, bifunctional Co(III)-salen catalyst for
alternating copolymerization of CO2 with cyclohexene oxide and terpolymerization with
aliphatic epoxides. Macromolecules 43:1396–1402
67. Nakano K, Kamada T, Nozaki K (2006) Selective formation of polycarbonate over cyclic
carbonate: copolymerization of epoxides with carbon dioxide catalyzed by a cobalt(III)
complex with a piperidinium end-capping arm. Angew Chem Int Ed 45(43):7274–7277
68. Scholl M, Ding S, Lee CW et al (1999) Synthesis and activity of a new generation of ruthenium-
based olefin metathesis catalysts coordinated with 1,3-dimesityl-4,5-dihydroimidazol-2-ylidene
ligands. Org Lett 1:953–956
69. Kim JG, Coates GW (2012) Synthesis and polymerization of norbornenyl-terminated
multiblock poly(cyclohexene carbonate)s: a consecutive ring-opening polymerization route
to multisegmented graft polycarbonates. Macromolecules 45(19):7878–7883
70. Geschwind J, Frey H (2013) Stable, hydroxyl functional polycarbonates with glycerol side
chains synthesized from CO2 and isopropylidene(glyceryl glycidyl ether). Macromol Rapid
Commun 34(2):150–155
71. Cohen CT, Chu T, Coates GW (2005) Cobalt catalysts for the alternating copolymerization of
propylene oxide and carbon dioxide: combining high activity and selectivity. J Am Chem Soc
127(31):10869–10878
72. Inoue S, Koinuma H, Tsuruta T (1969) Copolymerization of carbon dioxide and epoxide with
organometallic compounds. Makromol Chem 130:210–220
73. Inoue S, Koinuma H, Tsuruta T (1969) Copolymerization of carbon dioxide and epoxide.
J Polym Sci Part B Polym Lett 7(4):287–292
74. Aida T, Inoue S (1982) Synthesis of polyether-polycarbonate block copolymer from carbon
dioxide and epoxide using a metalloporphyrin catalyst system. Macromolecules 15:682–684
75. Lednor PW, Rol NC (1985) Copolymerization of propylene oxide with carbon dioxide: a
selective incorporation of propylene oxide into the polycarbonate chains, determined by
100 MHz carbon-13 NMR spectroscopy. Chem Commun (Cambridge, UK) 9:598–599
76. Nakano K, Hashimoto S, Nakamura M et al (2011) Stereocomplex of poly(propylene
carbonate): synthesis of stereogradient poly(propylene carbonate) by regio- and
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 197

enantioselective copolymerization of propylene oxide with carbon dioxide. Angew Chem Int
Ed 50(21):4868–4871
77. Byrnes MJ, Chisholm MH, Hadad CM et al (2004) Regioregular and regioirregular oligoether
carbonates: a 13C{1H} NMR investigation. Macromolecules 37:4139–4145
78. Salmeia KA, Vagin S, Anderson CE et al (2012) Poly(propylene carbonate): insight into the
microstructure and enantioselective ring-opening mechanism. Macromolecules
45:8604–8613
79. Salmeia KA (2012) Copolymerization of propylene oxide and CO2 with salen-type catalysts;
polymerization activities and polymer microstructure. Technische Universität München,
Munich
80. Chisholm MH, Navarro-Llobet D, Zhou Z (2002) Poly(propylene carbonate). 1. More about
poly(propylene carbonate) formed from the copolymerization of propylene oxide and
carbon dioxide employing a zinc glutarate catalyst. Macromolecules 35(17):6494–6504.
doi:10.1021/ma020348+
81. Ren W-M, Liu Y, Wu G-P et al (2011) Stereoregular polycarbonate synthesis: alternating
copolymerization of CO2 with aliphatic terminal epoxides catalyzed by multichiral cobalt
(III) complexes. J Polym Sci Part A Polym Chem 49:4894–4901
82. Nozaki K, Nakano K, Hiyama T (1999) Optically active polycarbonates: asymmetric alter-
nating copolymerization of cyclohexene oxide and carbon dioxide. J Am Chem Soc
121(47):11008–11009
83. Nakano K, Nozaki K, Hiyama T (2001) Spectral assignment of poly[cyclohexene oxide-alt-
carbon dioxide]. Macromolecules 34(18):6325–6332
84. Nakano K, Nozaki K, Hiyama T (2003) Asymmetric alternating copolymerization of
cyclohexene oxide and CO2 with dimeric zinc complexes. J Am Chem Soc 125:5501–5510
85. Cohen CT, Thomas CM, Peretti KL et al (2006) Copolymerization of cyclohexene oxide and
carbon dioxide using (salen)Co(III) complexes: synthesis and characterization of syndiotactic
poly(cyclohexene carbonate). Dalton Trans 1:237–249
86. Wu G-P, Ren W-M, Luo Y et al (2012) Enhanced asymmetric induction for the copolymer-
ization of CO2 and cyclohexene oxide with unsymmetric enantiopure salenCo(III) com-
plexes: synthesis of crystalline CO2-based polycarbonate. J Am Chem Soc 134:5682–5688
87. Stevens HC (1966) High-molecular-weight polycarbonates. US 324,841
88. Klaus S, Lehenmeier MW, Anderson CE et al (2011) Recent advances in CO2/epoxide
copolymerization-new strategies and cooperative mechanisms. Coord Chem Rev
255(13–14):1460–1479
89. Takeda N, Inoue S (1978) Polymerization of 1,2-epoxypropane and copolymerization
with carbon dioxide catalyzed by metalloporphyrins. Die Makromolekulare Chemie
179(5):1377–1381
90. Darensbourg DJ, Holtcamp MW, Struck GE et al (1999) Catalytic activity of a series of Zn
(II) phenoxides for the copolymerization of epoxides and carbon dioxide. J Am Chem Soc
121(1):107–116
91. Moore DR, Cheng M, Lobkovsky EB et al (2002) Electronic and steric effects on catalysts for
CO2/epoxide polymerization: subtle modifications resulting in superior activities. Angew
Chem Int Ed 41(14):2599–2602
92. Darensbourg DJ (2010) Chemistry of carbon dioxide relevant to its utilization: a personal
perspective. Inorg Chem (Washington, DC, USA) 49(23):10765–10780
93. Darensbourg DJ (2007) Making plastics from carbon dioxide: salen metal complexes as
catalysts for the production of polycarbonates from epoxides and CO2. Chem Rev
107(6):2388–2410
94. Kember MR, Buchard A, Williams CK (2011) Catalysts for CO2/epoxide copolymerization.
Chem Commun (Cambridge, UK) 47(1):141–163
95. Ikpo N, Flogeras JC, Kerton FM (2013) Aluminium coordination complexes in copolymer-
ization reactions of carbon dioxide and epoxides. Dalton Trans 42:8998–9006
198 P.T. Altenbuchner et al.

96. Sugimoto H, Inoue S (2004) Copolymerization of carbon dioxide and epoxide. J Polymr Sci
Part A Polym Chem 42(22):5561–5573
97. Coates GW, Moore DR (2004) Discrete metal-based catalysts for the copolymerization of
CO2 and epoxides: discovery, reactivity, optimization, and mechanism. Angew Chem Int Ed
43(48):6618–6639
98. Hansen KB, Leighton JL, Jacobsen EN (1996) On the mechanism of asymmetric nucleophilic
ring-opening of epoxides catalyzed by (Salen)CrIII complexes. J Am Chem Soc
118:10924–10925
99. Luinstra GA, Haas GR, Molnar F et al (2005) On the formation of aliphatic polycarbonates
from epoxides with chromium(III) and aluminum(III) metal-salen complexes. Chem Eur J
11:6298–6314
100. Darensbourg DJ, Yarbrough JC (2002) Mechanistic aspects of the copolymerization reaction
of carbon dioxide and epoxides, using a chiral salen chromium chloride catalyst. J Am Chem
Soc 124(22):6335–6342
101. Darensbourg DJ, Yarbrough JC, Ortiz C et al (2003) Comparative kinetic studies of the
copolymerization of cyclohexene oxide and propylene oxide with carbon dioxide in the
presence of chromium salen derivatives. In situ FTIR measurements of copolymer vs cyclic
carbonate production. J Am Chem Soc 125(25):7586–7591
102. Darensbourg DJ, Rodgers JL, Mackiewicz RM et al (2004) Probing the mechanistic aspects
of the chromium salen catalyzed carbon dioxide/epoxide copolymerization process using in
situ ATR/FTIR. Catal Today 98(4):485–492
103. Rao D-Y, Li B, Zhang R et al (2009) Binding of 4-(N, N-dimethylamino)pyridine to Salen-
and Salan-Cr(III) cations: a mechanistic understanding on the difference in their catalytic
activity for CO2/epoxide copolymerization. Inorg Chem (Washington, DC, USA)
48:2830–2836
104. Lu X-B, Wang Y (2004) Highly active, binary catalyst systems for the alternating copoly-
merization of CO2 and epoxides under mild conditions. Angew Chem Int Ed 43:3574–3577
105. Lu X-B, Shi L, Wang Y-M et al (2006) Design of highly active binary catalyst systems for
CO2/epoxide copolymerization: polymer selectivity, enantioselectivity, and stereochemistry
control. J Am Chem Soc 128(5):1664–1674
106. Cohen CT, Coates GW (2006) Alternating copolymerization of propylene oxide and carbon
dioxide with highly efficient and selective (salen)Co(III) catalysts: effect of ligand and
cocatalyst variation. J Polym Sci Part A Polym Chem 44(17):5182–5191
107. Niu Y, Zhang W, Pang X et al (2007) Alternating copolymerization of carbon dioxide and
propylene oxide catalyzed by (R, R)-SalenCoIII-(2,4-dinitrophenoxy) and Lewis-basic
cocatalyst. J Polym Sci Part A Polym Chem 45(22):5050–5056
108. Qin Z, Thomas CM, Lee S et al (2003) Cobalt-based complexes for the copolymerization of
propylene oxide and CO2: active and for polycarbonate synthesis. Angew Chem Int Ed
42(44):5484–5487
109. Sugimoto H, Kuroda K (2008) The cobalt porphyrin-lewis base system: a highly selective
catalyst for alternating copolymerization of CO2 and epoxide under mild conditions. Macro-
molecules 41:312–317
110. Guo L, Wang C, Zhao W et al (2009) Copolymerization of CO2 and cyclohexene oxide using
a lysine-based (salen)CrIIICl catalyst. Dalton Trans 27:5406–5410
111. Darensbourg DJ, Bottarelli P, Andreatta JR (2007) Inquiry into the formation of cyclic
carbonates during the (salen)CrX catalyzed CO2/cyclohexene oxide copolymerization pro-
cess in the presence of ionic initiators. Macromolecules 40(21):7727–7729
112. Darensbourg DJ, Phelps AL (2005) Effective, selective coupling of propylene oxide and
carbon dioxide to poly(propylene carbonate) using (salen)CrN3 catalysts. Inorg Chem
(Washington, DC, USA) 44(13):4622–4629
113. Darensbourg DJ, Mackiewicz RM (2005) Role of the cocatalyst in the copolymerization
of CO2 and cyclohexene oxide utilizing chromium salen complexes. J Am Chem Soc
127(40):14026–14038
7 Carbon Dioxide as C-1 Block for the Synthesis of Polycarbonates 199

114. Eberhardt R, Allmendinger M, Rieger B (2003) DMAP [4-(N, N-dimethylamino)pyridine]/


Cr(III) catalyst ratio: the decisive factor for poly(propylene carbonate) formation in the
coupling of CO2 and propylene oxide. Macromol Rapid Commun 24(2):194–196
115. Darensbourg DJ, Mackiewicz RM, Rodgers JL et al (2004) Cyclohexene oxide/CO2 copoly-
merization catalyzed by chromium(III) salen complexes and N-methylimidazole: effects of
varying salen ligand substituents and relative cocatalyst loading. Inorg Chem (Washington,
DC, USA) 43:6024–6034
116. Darensbourg DJ, Mackiewicz RM, Rodgers JL et al (2004) (Salen)CrIIIX catalysts for the
copolymerization of carbon dioxide and epoxides: role of the initiator and cocatalyst. Inorg
Chem (Washington, DC, USA) 43:1831–1833
117. Moore DR, Cheng M, Lobkovsky EB et al (2003) Mechanism of the alternating copolymer-
ization of epoxides and CO2 using β-diiminate zinc catalysts: evidence for a bimetallic
epoxide enchainment. J Am Chem Soc 125(39):11911–11924
118. Cyriac A, Lee SH, Varghese JK et al (2010) Immortal CO2/propylene oxide copolymeriza-
tion: precise control of molecular weight and architecture of various block copolymers.
Macromolecules 43(18):7398–7401
119. Noh EK, Na SJ, Sujith S et al (2007) Two components in a molecule: highly efficient and
thermally robust catalytic system for CO2/epoxide copolymerization. J Am Chem Soc 129
(26):8082–8083
120. Yoo J, Na SJ, Park HC et al (2010) Anion variation on a cobalt(III) complex of salen-type
ligand tethered by four quaternary ammonium salts for CO2/epoxide copolymerization.
Dalton Trans 39(10):2622–2630
121. Na Sung J, Sujith S, Cyriac A et al (2009) Elucidation of the structure of a highly active
catalytic system for CO2/epoxide copolymerization: a salen-cobaltate complex of an unusual
binding mode. Inorg Chem (Washington, DC, USA) 48(21):10455–10465
122. Ren W-M, Liu Z-W, Wen Y-Q et al (2009) Mechanistic aspects of the copolymerization of
CO2 with epoxides using a thermally stable single-site cobalt(III) catalyst. J Am Chem Soc
131(32):11509–11518
123. Zheng YQ, Lin JL, Zhang HL (2000) Crystal structure of zinc glutarate, Zn(C5H6O4).
Zeitschrift fuer Kristallographie – New Crystal Struct 215(4):535–536
124. Ree M, Hwang Y, Kim J-S et al (2006) New findings in the catalytic activity of zinc glutarate
and its application in the chemical fixation of CO2 into polycarbonates and their derivatives.
Catal Today 115:134–145
125. Eberhardt R, Allmendinger M, Zintl M et al (2004) New zinc dicarboxylate catalysts for the
CO2/propylene oxide copolymerization reaction: activity enhancement through zn(II)-
ethylsulfinate initiating groups. Macromol Chem Phys 205(1):42–47
126. Zhu Q, Meng YZ, Tjong SC et al (2003) Catalytic synthesis and characterization of an
alternating copolymer from carbon dioxide and propylene oxide using zinc pimelite. Polymer
Int 52:799–804
127. Wang JT, Shu D, Xiao M et al (2006) Copolymerization of carbon dioxide and propylene
oxide using zinc adipate as catalyst. J Appl Polym Sci 99:200–206
128. Klaus S, Lehenmeier MW, Herdtweck E et al (2011) Mechanistic insights into heterogeneous
zinc dicarboxylates and theoretical considerations for CO2-epoxide copolymerization. J Am
Chem Soc 133:13151–13161
129. Lehenmeier M (2012) Synthese zweikerniger Zinkkatalysatoren zur Copolymerisation von
Kohlenstoffdioxid und Epoxiden. Technische Universität München, Munich
130. Van Meerendonk WJ, Duchateau R, Koning CE et al (2005) Unexpected side reactions and
chain transfer for zinc-catalyzed copolymerization of cyclohexene oxide and carbon dioxide.
Macromolecules 38(17):7306–7313
131. Allen SD, Moore DR, Lobkovsky EB et al (2002) High-activity, single-site catalysts for
the alternating copolymerization of CO2 and propylene oxide. J Am Chem Soc
124(48):14284–14285
200 P.T. Altenbuchner et al.

132. Rieth LR, Moore DR, Lobkovsky EB et al (2002) Single-site β-diiminate zinc catalysts for
the ring-opening polymerization of β-butyrolactone and β-valerolactone to poly
(3-hydroxyalkanoates). J Am Chem Soc 124:15239–15248
133. Eberhardt R, Allmendinger M, Luinstra GA et al (2003) The ethylsulfinate ligand: a highly
efficient initiating group for the zinc β-diiminate catalyzed copolymerization reaction of CO2
and epoxides. Organometallics 22:211–214
134. Lee BY, Kwon HY, Lee SY et al (2005) Bimetallic anilido-aldimine zinc complexes for
epoxide/CO2 copolymerization. J Am Chem Soc 127(9):3031–3037
135. Bok T, Yun H, Lee BY (2006) Bimetallic fluorine-substituted anilido-aldimine zinc com-
plexes for CO2/(cyclohexene oxide) copolymerization. Inorg Chem (Washington, DC, USA)
45(10):4228–4237
136. Jutz F, Buchard A, Kember MR et al (2011) Mechanistic investigation and reaction kinetics
of the low-pressure copolymerization of cyclohexene oxide and carbon dioxide catalyzed by
a dizinc complex. J Am Chem Soc 133(43):17395–17405
137. Kember MR, White AJP, Williams CK (2010) Highly active di- and trimetallic cobalt
catalysts for the copolymerization of CHO and CO2 at atmospheric pressure. Macromole-
cules 43:2291–2298
138. Buchard A, Kember MR, Sandeman KG et al (2011) A bimetallic iron(III) catalyst for CO2/
epoxide coupling. Chem Commun (Cambridge, UK) 47:212–214
139. Kember MR, Williams CK (2012) Efficient magnesium catalysts for the copolymerization of
epoxides and CO2; using water to synthesize polycarbonate polyols. J Am Chem Soc 134
(38):15676–15679
140. Lehenmeier MW, Kissling S, Altenbuchner PT et al (2013) Flexibly tethered dinuclear zinc
complexes: a solution to the entropy problem in CO2/epoxide copolymerization catalysis?
Angew Chem Int Ed 52(37):9821–9826
Part II
Photocatalytic, Electrochemical
and Inorganic Reactions
Chapter 8
Fixation of Carbon Dioxide Using Molecular
Reactions on Flexible Substrates

Jacob Jensen and Frederik C. Krebs

8.1 Introduction

The largest chemical challenge in terms of scale is undoubtedly the control and
management of carbon dioxide in the atmosphere. Due to annual man-made
emissions of 24 gigaton CO2, the atmospheric content of CO2 has increased from
270 to 385 ppm. Totally this amounts to around 1 teraton excess carbon dioxide in
the atmosphere stemming mainly from the combustion of fossil fuels [1]. This is a
result of a very efficient and a highly distributed system for conversion of reduced
forms of carbon into carbon dioxide (cars, households, etc.). The increased CO2
generated by mankind is not balanced by an equal amount of fixation mainly due to
deforestation. There is broad recognition worldwide that the emission of carbon
dioxide must be brought down and preferably reverted such that a situation of a
negative man-made carbon dioxide emission is reached. Currently the scale of that
challenge exceeds our capacity for handling materials by several orders of magni-
tude, and developing solutions to this should be considered as a priority.
Laboratory-scale catalysts are therefore not particularly useful unless they can be
scaled to the required level without any carbon footprint. In addition to this
problem, it is impossible to extract carbon dioxide directly from the atmosphere
without generating more carbon dioxide in the process (i.e., the sequestration of
1 kg of carbon dioxide from the atmosphere will emit more than 1 kg of carbon
dioxide). This fundamental problem has resulted in the passive view that man-made
carbon dioxide emissions should be reduced but not reverted. This is done most
efficiently by reducing carbon dioxide emissions directly at the source, e.g., flue
gasses from coal fired power plants, iron smelters, and concrete factories.
Extracting the carbon dioxide from the flue gasses is costly and the carbon seques-
tration and storage (CSS) costs can be as high as 50 % of the electricity costs.

J. Jensen • F.C. Krebs (*)


Department of Energy Conversion and Storage, Technical University of Denmark,
Frederiksborgvej 399, DK-4000 Roskilde, Denmark
e-mail: frkr@dtu.dk

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 203
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_8,
© Springer-Verlag Berlin Heidelberg 2014
204 J. Jensen and F.C. Krebs

In order to propose an efficient solution to the problem, one must use a renewable
energy source to convert atmospheric carbon dioxide either for reuse of the stored
energy or by fixation of it, thus reducing the amount of carbon dioxide in the
atmosphere. The solar energy available at the surface of the Earth is around
1,000 W·m2 (AM1.5G), and depending on the local climate, one can expect to
receive between 1,000 and 2,000 kWh m2 year1 from solar irradiation. Based on
this, a rational approach seems to be one where sunlight is used to drive carbon
dioxide converting reactions. In order to efficiently address the issue of scale, the
development of chemical conversion systems must reflect this. A solution to this
challenge is through the use of flexible substrates as carrier foils for the conversion
agents. By employing flexible substrates, the large areas necessary for solar expo-
sure conversion can be achieved through fast coating of low-cost photocatalysts or
fixation agents using roll-to-roll processing at low temperatures. The following
chapter will address the challenges encountered in utilizing this strategy.

8.2 Thermodynamical Considerations

The carbon dioxide molecule is linear and has two polar bonds between the
electrophilic carbon and the nucleophilic oxygen atoms. CO2 is a rather unreactive
molecule, owing to the fact that it is in its most oxidized state. This is mirrored in
the large standard enthalpy of formation at 298 K which is 393 kJ·mol1
(reaction 8.1).

C þ O2 ! O¼C¼O ð8:1Þ

The stability of the CO2 molecule requires that a large amount of energy is
supplied in order for transformations to occur. Depending on the product, the
energy requirements for reduction of CO2 are as follows for carbon monoxide
and methane:

CO2 ! CO þ ½O2 ΔH ¼ 283 kJ  mol1 ; ΔG ¼ 257 kJ  mol1 ð8:2Þ


 1
CO2 þ H2 O ! CH4 þ 2O2 ΔH ¼ 285 kJ  mol ;

ΔG ¼ 237 kJ  mol1 ð8:3Þ

For the Gibbs free energy to be zero, reaction (8.2) needs to be carried out at a
minimum of 3,075  C, if the process was to be accomplished thermochemically
[2]. For this to be energetically viable, the energy should come from a renewable
source, such as concentrated solar irradiation [3]. Such high temperatures are not
compatible with the flexile substrate approach described in this chapter. For
that approach to be useful, only low temperature (<100  C) can be employed,
and photocatalysis using transition metal oxide semiconductors is one way of
achieving this.
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 205

Fig. 8.1 The energy levels necessary for photocatalytic reduction. (a) Upon solar irradiation an
electron is excited from the valence band (VB) to the conducting band (CB), from where it is accessible
for reduction of an acceptor molecule (e.g., CO2). The excitation creates an electron deficiency or
hole in semiconductor. Simultaneously a donor molecule (e.g., water) donates an electron to the hole,
completing the catalytic circle. (b) The visible part of the electromagnetic spectrum, showing
the wavelength below and photon energies above. The arrows indicate the reduction potential of
CO2, the optimal band gap for practical purposes, and the band gap of TiO2, respectively

8.2.1 Band Gap and Photocatalytic Properties

The basic process in photocatalysis is depicted in Fig. 8.1a. As semiconductors lack


coherent energy levels, there is a void region between the valence band (VB) and the
conducting band (CB), where energy transitions are not allowed, known as the band
gap. If photons of energy equal to or higher than the band gap reaches the semicon-
ductor surface, an electron is excited to the conducting band leaving behind a
positively charged electron vacancy or hole (h+). The energy levels at the bottom of
the semiconductor conducting band is the reduction potential of the excited electron,
while the top of the valence band is the oxidation potential of the holes generated
[4]. The catalytic property of the semiconductor is the ability to simultaneously reduce
one adsorbed species while oxidizing another. For reduction to occur, the reduction
potential of the adsorbed species needs to be higher (more positive) than the conduc-
tion band of the photocatalyst, while oxidation requires the reduction potential (of the
adsorbed species) be lower (more negative) than the energy of the valence band holes;
this is shown in Fig. 8.1a. For carbon dioxide conversion, CO2 is the acceptor
molecule, while compounds such as hydrogen or water acts as donor molecules.
If reaction (8.2) is to occur by photocatalysis, the band gap of the photocatalyst
should be at least 1.33 eV (per photon), which corresponds to wavelengths below
930 nm (E(eV) ¼ 1,240 eV·λ(nm)). But as entropy changes and other losses
associated with the splitting of CO2 have to be taken into account, the optimal
energy required per photon is in the region from 2 to 2.4 eV, see Fig. 8.1b [5]. Of the
semiconductors commonly used, titanium dioxide is by far the most well studied.
TiO2 is found in several crystal forms, with anatase and rutile being the forms
206 J. Jensen and F.C. Krebs

usually employed for photocatalysis. Due to differences in lattice structures and


electronic densities, anatase and rutile have different band gaps of 3.20 eV and
3.02 eV, respectively. This corresponds to absorption thresholds of 384 and 410 nm
[4]. It is still unclear whether or not the band edges of TiO2 is such that
photocatalysis can occur without doping the catalyst, but it is our experience that
photocatalytic reactions using unmodified TiO2 are possible [4, 6–10].

8.3 Carbon Dioxide Conversion Reactors

In order to convert CO2 into useful products, a suitable reactor needs to be developed.
Ideally the source of CO2 would come from the atmosphere, but since the amount of
CO2 in the atmosphere is approximately 0.04 %, the available CO2 for conversion is
low. Moreover, the conversion system risks being polluted by other atmospheric
components such as volatile sulfur compounds. All in all purification of CO2 must be
accomplished before conversion. The carbon capture and sequestration technologies
commonly used for CO2 capture today are characterized by high energy consumption
for desorption of CO2, which makes using these technologies for purification of CO2
in connection with photocatalytic reduction questionable. However, if efficient
capture system like CO2 diffusion selective membranes were employed, extraction
from the atmosphere would be possible using a minimum amount of energy
[11, 12]. Another solution is to use the Direct Air Capture (DAC) system, where
air is cycled through solid alkali sorbents that binds the weakly acidic CO2. The
concentrated CO2 becomes available by replacement by water at low temperatures
(40–45  C) [13]. By combining a membrane system with an appropriate
photocatalyst or fixation reagent (roll-to-roll coated onto the polymer membrane),
one can manufacture a large area foil that extracts CO2 from the atmosphere and
converts it to an energy-rich chemical form (e.g., a liquid fuel) when exposed to
sunlight. A schematic for such a unit is shown in Fig. 8.2. In Fig. 8.2a, the conversion
agent is coated onto a membrane that selectively allows CO2 to permeate into the
reaction chamber. The other side of the chamber consists of a window that allows
sunlight to pass. The energy for conversion is supplied by solar irradiation, and
products are recovered as liquid/gaseous energy carriers or as fixated precursors.
A slightly different conversion unit is depicted in Fig. 8.2b, where CO2 in a similar
way is separated from the atmosphere by a membrane. The difference in this setup is
that the conversion reagent is coated onto a flexible substrate that also functions as the
reactor window.
The conversion unit suggested is a complex chemical system, and several issues
must be solved in order for successful operation:
– Coating of conversion reagent
– Stability of window substrate
– Reactivity of conversion reagent
– Analysis of products
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 207

Fig. 8.2 Two examples of CO2 conversion units that function by extraction of atmospheric CO2
using a selective membrane. In (a) the conversion reagent is coated onto the membrane, whereas in
(b) it is coated onto a flexible substrate that functions as the reactor window

To address these questions most efficiently, simple reactor designs that focus on
one or two aspects are employed. An example of such a prototype is depicted in
Fig. 8.3, which is comparable to similar reactors [10, 14, 15]. This reactor is made
of a lathed aluminum chamber with an inner volume of 290 cm3. In order to gain
access to the chamber, changing of the front window, cleaning of the chamber, etc.,
the demountable top is removed by unscrewing, and it is tightly closed again by
screwing the window against an o-ring gasket in the housing (see Fig. 8.3b, c).
208 J. Jensen and F.C. Krebs

Fig. 8.3 (a) Schematic of simple prototype conversion unit. (b) An operational conversion reactor
equipped with a quartz window. (c) PET foil to be used as reactor window. (d) Alumina inserts
allowing variations in reactor volume

Introduction of gaseous components and flushing of the chamber are done through
of Swagelock™ valves. Variation of the chamber volume is possible as blocks of
aluminum were lathed to fit the chamber (see Fig. 8.3d). The chamber construction
allows for several modes of usage as discussed previously (see schematics in
Fig. 8.2). By placing the coated substrate inside the reactor facing out and using a
window material that allows passage of solar irradiation, conditions like those in
Fig. 8.2a are established. By choosing the appropriate material, the window can
function as a solar filter, which is beneficial when optimizing operational parame-
ters. A more unconventional operational mode, like that schematized in Fig. 8.2b, is
to coat the conversion reagent onto the front window, which can be a flexible
substrate like PET (Fig. 8.3c) or quarts (Fig. 8.3b).
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 209

8.3.1 Infrared Spectroscopy

The method of analysis is another thing that can successfully be integrated in the
reactor design. In photocatalytic conversion reactors like the one in Fig. 8.3b, the
products are generally gasses, which can be detected by either GC-MS or infrared
spectroscopy [16–18]. The reactor in Fig. 8.3b is equipped with zinc-selenide
(ZnSe) windows and makes use of in situ infrared spectroscopy for analysis. The
setup is such that the reactor can be removed from the light source during operation.
After IR analysis (<1 min), the reactor is placed back under the light source and
irradiation is resumed. The use of in situ infrared spectroscopy was found partic-
ularly useful, being a nondestructive method that combines chemical fingerprinting
with a reasonable sensitivity that enables detection of many different materials. As
the products of photocatalytic CO2 conversion are simple small molecules like CO,
CH4, and CH3OH, the identification by IR is straightforward and can be a quanti-
tative analytical tool by making standard calibration curves of components such as
methane, formic acid, methanol, and carbon monoxide (the detection limit of CO
using the reactor in Fig. 8.2b is around 0.01 %). Table 8.1 shows the major bands of
some simple carbon volatiles [18–23].
Using isotopically labeled 13CO2 is a particular useful way of distinguishing
between reaction products formed from introduced carbon dioxide or carbon
dioxide originating from sources such as the flexible substrate, surfactants, and
catalyst precursor [25, 26]. The difference in the IR spectra of reduction products
containing 13C is determined using the harmonic equation [25].
In the case of carbon monoxide, the difference in IR spectra is shown in Fig. 8.4.
Spectrum (A) displays the infrared spectrum of 12CO, where the C–O stretching is
centered at 2,143 cm1. This vibration has moved to 2,096 cm1 in spectrum (B)
that displays 13CO. Spectrum C is a mixture of the two isotopes [10, 18, 20].

8.3.2 Adsorption of Carbon-Containing Volatiles

When designing reactor conversion reactors, one naturally has to think about the
materials used. Using all quartz reactors is possible for small-scale reactors, but
scaling these quickly becomes unsolvable, which is why most reactors consist of a
chamber comprising a window. What might not be so obvious is that the gaseous
components have the ability to adsorb onto the chamber surface and herby compli-
cate the analysis of the products, when desorbing at a later stage. This was observed
for the chamber in Fig. 8.3, where isotopically labeled 13CO2 and 12CO was
detected in the reaction chamber devoid of any catalyst and after prolonged flushing
with pure hydrogen gas and subsequent UV illumination. As a precaution, a
flushing procedure aimed at desorbing CO2 from the chamber surface prior to
operation was installed. This consisted of heating the chamber to 65  C for
60 min, flushing with hydrogen, and evaporation at reduced pressure; a procedure
210 J. Jensen and F.C. Krebs

Table 8.1 Major infrared bands for simple carbon volatiles. For additional
spectral information, the reader is directed to the National Institute of
Standards and Technology’s Chemistry Webbook [24]
Compound Major bands (cm1)
CO2 2,349
CO 2,143
CHO 2,850, 2,785, 1,750, 1,485, 1,250
CH3OH 3,337, 2,934, 2,822, 1,029, 655
CH4 3,020, 1,534, 1,305

Fig. 8.4 Infrared absorbance spectra of gaseous carbon monoxide isotopes in argon (1 cm1
resolution). (a) 12CO. (b) 13CO. (c) A mixture of 12CO and 13CO

that was repeated twice. 13CO2 was believed to originate from earlier experiments
performed in the reaction chamber as the aluminum surface was subjected to
water vapor, oxygen, and light. Reaction conditions like these lead to the formation
of alumina and aluminum hydroxide, which adsorbs volatiles such as CO2 and CO
[14, 27]; properties that has been exploited in carbon dioxide capture and seques-
tration processes. In addition, it has also been shown that a range of carbon-oxygen-
containing species possibly resides on the alumina surface, including carbon
monoxide that binds to Lewis acid sites on alumina, although to a lesser degree
than carbon dioxide [28–30]. The complete desorption of carbon dioxide from
alumina has been shown to take place at around 200  C for mesoporous alumina
and γ-alumina [31–33].
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 211

8.4 Coating

While numerous methods exist for depositing of material onto a substrate, energy
considerations has to be acknowledged when developing devices that convert
carbon dioxide. Employing energy-demanding deposition methods like sputtering,
vacuum deposition, or thermal evaporation requires that the conversion agent is
equally effective in order to compensate for the excessive energy used. Naturally all
deposition methods require energy, but in order to avoid high CO2 conversion
requirements, the following section describes some of the deposition methods
available that minimizes the impact on the energy balance. Common to these
techniques are that a wet film of the dissolved material is formed upon deposi-
tion/coating, followed by evaporation of the solvent whereupon a thin film is
formed. For laboratory scale, two established methods are spin- and spray coating,
which are both facile with simple requirements to equipment. For spray coating,
only a spray gun and a connection to pressurized gas are required (see Fig. 8.5).
More advanced forms of this coating method are available, e.g., ultrasonic spray
coating, thermal spray coating, and cold spray coating. In spray coating the
dissolved photocatalyst is supplied to a nozzle where tiny aerosol droplets are
formed when an inert gas (usually nitrogen or argon) is supplied. When vaporized
the increased area of the droplets will cause partly evaporation of the solvent prior
to impact. On impact the droplets are smashed and following the remaining solvent
evaporate leaving behind a thin film of the desired material. Manually spray coating
takes some practice as it is difficult to get an even film and it is preferable to use a
solvent with a low vapor pressure, since this will increase the speed of evaporation.
Solvents with high vapor pressures (e.g., chlorobenzene, dichlorobenzene) tend to
produce uneven films due to streaks of runny solvents. Spray coating yields thin
films with a rough surface, as it consists of clusters of materials (not necessarily in
contact with other clusters). The surface roughness makes thickness determination
troublesome, but the advantage of the roughness is an increased surface area that is
beneficial in terms of photocatalytic activity.
Spray coating can be automatized and is in principle compatible with roll-to-roll
coating. When spray coating it is preferable that the photocatalyst is dissolved prior

Fig. 8.5 (a) Schematic of a simple spray gun. The solution to be sprayed is supplied from a liquid
reservoir and extruded through the nozzle as an aerosol by pressurized gas (typically N2 or Ar).
(b) Picture of a simple spray gun. This type has the reservoir directly attached to the gun
212 J. Jensen and F.C. Krebs

Fig. 8.6 (a) Schematic of the spin coating process. The ink to be spun cast is placed on the
substrate prior to spinning and supplied while the substrate is in motion. By centrifugal forces, the
ink is forced from the center of the disc to the edges. Evaporation of the solvent produces a thin
solid film (Reproduced from Ref. [34] by permission of The Royal Society of Chemistry).
(b) Picture of a spin coater. The substrate is put onto the rotating disc and held in place by vacuum
while spinning. The panel on the left controls the rotational velocity and enables basic program-
ming of the spin coater

to spraying, as the use of a slurry of precipitated photocatalyst will require a very


low “concentration” equivalent of an increased amount of solvent.
Spin coating is another well-known technique that is fast, easy to use, and
requires a minimum of equipment (see Fig. 8.6) [34]. An advantage of spin coating
is that the results are highly reproducible, yielding films of identical thickness and
morphology under identical conditions. These advantages make spin coating an
excellent technique for small-scale laboratory experiments. The disadvantage is the
large amount of material is wasted in the process, which makes this method very
cost-inefficient and unsuitable for larger-scale production. In addition manually
changing and cleaning substrates are time consuming.
While spin and spray coating are well suited for small-scale lab experiments, the
disadvantages due to the large amount of manual work make them unsuited for
larger-scale manufacturing tasks. Slot-die coating is ideally suited for this task as it
is a fast and robust technique that is roll-to-roll compatible. In slot-die coating, the
ink is dispersed onto a substrate through a coating head by a pump (see Fig. 8.7).
The substrate is carried forward by a rotating drum. After solvent evaporation, a
thin film is formed. The thickness of the film is controlled by ink concentration,
rotational speed of the drum, and the speed of deposition. Laboratory-scale roll
coaters enable limits on the amount of ink necessary for coating and are useful for
optimizing procedures for larger-scale roll-to-roll equipment [35, 36].
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 213

Fig. 8.7 (a) Schematic of the slot-die coating technique showing the coating head above the
substrate. Following the inlet the ink is introduced into a small reservoir, where from it flows
through void and is dispersed onto the substrate (Reprinted from Ref. [35], Copyright 2009, with
permission from Elsevier). The substrate is carried forward by the rotating drum. (b) A picture of a
slot-die coating head (top) with a tube attached to the ink inlet. Part of the rotating drum is seen at
the bottom of the picture (no substrate attached)

8.4.1 Surfactants and Carbonaceous Residues

Inorganic photocatalysts usually require organic surfactants in order to meet the


solubility requirement for the film-forming techniques described above. In addition
to dissolution of the inorganic particles, the surfactants ensure the correct growth
of the metal oxide nanoparticles and suppress aggregation, agglomeration, and
precipitation [37–44]. The list of surfactants is extensive, and a few examples are
given in Fig. 8.8.
Since the photocatalytic activity of a catalyst depend on the proper crystal phase of
the catalyst, the surfactant group needs to be removed after coating. Of particular
interest when working with flexible substrates is the removal of surfactants at low
temperatures, as this allows for the exposure of the catalytic surface to reactants
without thermal degradation of the polymer substrate. This can be done by exploiting
the photooxidative abilities of TiO2, whereby, e.g., oleic acid is decomposed to
alkanes [45–49].
An example of photocatalytic removal of oleic acid on TiO2 is given in Fig. 8.9.
The graph shows the infrared spectra of TiO2 nanoparticles in a KBr pellet during
irradiation (1,200 W·m1). The characteristic bands of oleic acid capped TiO2 at
2,924, 2,852, 1,523, and 1,433 cm1 are shown in spectrum A. Spectra B, C, and D
shows the IR spectra after 60, 120, and 360 min of UV light, respectively. As seen
from Fig. 8.5, the removal of oleic acid by UV light at low temperatures is a viable
214 J. Jensen and F.C. Krebs

Fig. 8.8 Surfactants used a O


for stabilization of
photocatalyst nanoparticles. OH
(a) Oleic acid, (b) stearic
acid, (c) sodium dodecyl
O
sulfate, (d) primary alkyl b
amines, (e) polyethylene OH
glycol

c
O SO3 Na

d e
NH2 O OH
n H n

Fig. 8.9 Infrared analysis of the removal of oleic acid by UV-light. (a) Oleic acid capped TiO2
particles prior to irradiation. (b–d ) Sample following 60, 120, and 360 min of irradiation,
respectively

process. The disappearance of the characteristic bands after a relatively short while
(~60 min) leaves a broad band centered at 3,433 cm1 and one at 1,626 cm1. The
former is attributed to water adsorbed on titanium or hydroxyl groups; the latter is
attributed to water adsorbed onto KBr. The presence of these bands is not
disturbing, since water is introduced in the reaction chamber prior to the photo-
chemical reaction [50, 51].
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 215

Fig. 8.10 Roll of flexible


PET foil

Whether thermal or photocatalytic procedures are employed to decompose the


surfactant group, it is important to realize the presence of carbonaceous residues on the
catalyst surface following surfactant decomposition [52]. For oleic acid thermally
decomposed in air, infrared analysis show absorption bands in the carbonyl region
(2,200, 2,259 and 2,330 cm1), suggesting that the leftover residues are COx com-
plexes [53]. In addition to the oxygen-containing species, amorphous carbon/graphite
residues are formed when the decomposition takes place in an argon atmosphere, and
alkane residues are believed to adsorb on the metal oxide surface beneath this graphite
shell. The presence of carbon-containing species should be noted as they contribute to
the complexity in analyzing the product mixture.

8.4.2 Flexible Substrates as Carrier Foils

By employing roll-to-roll coating of the photocatalyst onto flexible substrates, large


area photocatalytic surfaces can be achieved at high production rates. The choice of
substrate is wide as long as thin flexible foils are possible (see Fig. 8.10). Standard
choices for roll-to-roll substrates are foils made of poly(ethylene terephthalate)
(PET), poly(ethylene napthalate) (PEN), and poly(ethylene) (PE); the structures of
these three polymer substrates are depicted in Fig. 8.11.
While employing flexible substrates is advantageous for large-scale manufacture
and processing, there are questions that need to be addressed in order to use these as
carrier foils for photochemical conversion of carbon dioxide. An obvious challenge
is that polymeric films degrade when exposed to solar irradiation, oxygen, and
moisture, i.e., conditions that are found outdoors where the photocatalytic system
216 J. Jensen and F.C. Krebs

O O O
O H H
O O
C O O
C C
H H

PET PEN PE

Fig. 8.11 Structure of poly(ethylene terephthalate) (PET), poly(ethylene napthalate) (PEN), and
poly(ethylene) (PE). The figure shows the repeating unit of the polymeric structure

are to operate [54–59]. Moreover, as the oxidative degradation products include


volatiles such as carbon dioxide and carbon monoxide, employing polymeric sub-
strates for photoconversion of CO2 could seem like Sisyphus’ work [60–63]. In
spite of this, polymeric substrates foils remain interesting as carrier foils since it is
the total amount of carbon dioxide converted that matter, i.e., the amount of carbon
dioxide photocatalytically converted or bound should be larger than the amount of
oxidized carbon species released during substrate decomposition. Two other factors
support this: (1) Coating of absorbing nanoparticle can in some instances protect
the substrate material from solar irradiation. (2) By construction of a system that
eliminates access to atmospheric oxygen. To assess the viability of polymeric
substrates and aid in designing reactors for such, a short review of the fundamental
degradation processes is appropriate.
It has been found that photodegradation of polymer substrates is a combination of
two reactions: photolysis and photocatalytic oxidation [60, 63–70]. Photolysis is the
“direct” cleavage, whereby photoinitiated radicals generated in the polymeric sub-
strate lead to cleavage by, e.g., a Norrish type I reaction pathway as depicted in
Scheme 8.2. Photooxidation is likewise initiated by radicals generated in the polymer
but differs in that addition of oxygen generate hydroperoxide species that decompose
as shown in Schemes 8.1 and 8.3 for PE and PET, respectively. Since photooxidation
involves molecular oxygen, it should be possible to negate this reaction by employing
an oxygen-free atmosphere. However, oxidation of the substrate has been found to
occur even though oxygen-free conditions were employed [14]. This puzzling obser-
vation was attributed to the reduction of TiO2 or residual substrate oxygen or carbon
dioxide but do in any instance testify to the complexity of these chemical systems and
the care that must be taken when evaluating results hereof.
Due to the difference in chemical composition of polyolefins (PE, PP, etc.) and
polyesters (PET, PBT, etc.), the predominant degradation mechanism differs. Though
multiple pathways might occur simultaneously, the predominant reaction for irradi-
ated polyolefins in an oxygen-containing atmosphere is photocatalytic oxidation.
This is due to the absence of chromophoric groups capable of absorbing energy
from terrestrial sunlight (>290 nm), which result in insufficient energy for direct
bond cleavage. However, light absorbing chromophores are found to develop in the
films during processing and storage of polyolefins, which make generation of active
intermediates on the surface possible. These react with oxygen and decompose as
shown in Scheme 8.1 [66]. Since PET and related polyesters contain chromophoric
groups as part of their molecular structure (aromatic and ester groups), they are
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 217

O OH OH
UV light O2 O O
H2 C

O O O2 O R-H
R-H / O2 C C C H2C CO2
R HO O O

CO

Scheme 8.1 Photooxidation of polyolefins. Reactive intermediates are generated on the surface of
the substrate by UV irradiation. This is followed by addition of oxygen and proton abstraction to
yield hydroperoxide species that decompose to carbonyl compounds and alkyl radicals. The alkyl
radicals react further with oxygen to generate carbonyl radicals that decompose to CO or CO2 via
another sequence of oxygen addition and proton abstraction

O O O O
O O OCHCH2
C C O C O
CO
Ph
O
O
O
C O
CH2CH2

O O H CO2
Ph
C CHO O O
O CH2 O
C O

HOOCH2Ph
Norrish type I pathways

O OH
C H2C O
O

Norrish type II pathway

Scheme 8.2 Photolytic pathways for PET and related polyester substrates

O OH O
O O O2/R-H O O O O
O O O
C O C O C O

O O O O
O
C O O O
C CO2 HO

Scheme 8.3 Photocatalytic oxidation mechanism of PET and related polyesters


218 J. Jensen and F.C. Krebs

Fig. 8.12 Effect of TiO2 nanoparticles on flexible substrates during irradiation. When flexible
substrates are subjected to irradiation (top), the coated particles can affect the substrate in two
ways. (a) Protect the substrate by scattering the irradiation or absorb irradiation without electron
excitation. (b) Enhance photooxidation whereby the substrate donates electrons to the acceptor
(designated A) via the photocatalyst

capable of absorbing high energy irradiation, which make them susceptible to direct
photolytic cleavage that can proceed via Norrish type I or II pathways as illustrated in
Scheme 8.2 [58, 60, 63–65].
While the photolytic cleavage is the predominant mechanism in PET and related
polyesters, photooxidative decomposition is also observed analogous to that seen
for polyolefins. Radical species, capable of reaction with oxygen, are produced in
the surface by irradiation. This yields hydroperoxide species that decompose to
various products including carbon dioxide (see Scheme 8.3) [14, 63, 71, 72].

8.4.3 Effect of TiO2 Coatings

Chemical species on the substrate surface has been found to alter the photo and
thermal stability, albeit the results of such coatings are ambiguous. On one hand,
TiO2 can attenuate decomposition and protect the substrate by scattering and
absorbing UV irradiation. On the other hand, TiO2 has been exploited in environ-
mental management such as wastewater treatment and self-cleaning surfaces where
it functions by photooxidizing organic molecules – a process that could enhance
degradation in the substrate [45, 73–75]. The former situation is depicted in
Fig. 8.12a, and the latter in Fig. 8.12b.
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 219

The effect of TiO2 has been studied in some detail for paint (where TiO2 is used as
a filler/pigment) and films containing TiO2. Results from these studies show that the
effect depends on the crystal form of the oxide. Rutile TiO2 contains stable OH
groups on the surface and hence stabilize the substrate, whereas the anatase form of
TiO2 is reactive and consequently enhance photooxidation [14, 76–78]. Furthermore,
commercially available foils contain various stabilizers, additives, and surfactants
that could interact with TiO2 and either enhance photooxidation or protection. In
order to successfully convert small-scale laboratory experiments into large-scale
photocatalytic systems, the atmosphere composition must be taken into account.
Prior to operation, the amount of evolved CO2 must be measured if flexible substrates
are used as carrier foils. This is best done under CO2-free conditions to be as reliable
as possible, but at the same time the atmosphere should mimic the operational
conditions as closely as possible. The operational conditions depend on the reactor
and the concentration of CO2 employed. But factors such as temperature and pressure
are also of importance as described.

8.5 Conclusion and Perspectives

Photocatalytic reduction of CO2 is a promising way of fighting the increasing


content of atmospheric carbon dioxide, but as this technology is still being explored
at laboratory scale, chances are that it takes several decades for efficient conversion
systems to appear. For such a system to be effective, issues must be addressed that
all relate to the enormous scale of the challenge.
The mechanism of photocatalytic reduction is not completely solved yet, even
though qualified assumptions are plentiful. This leads to rough predictions on
product composition, but most experiments seems to be on trial and error basis.
As this chapter illustrates, monitoring CO2 conversion is difficult. The adsorption-
desorption of CO2 from reactor surfaces, porous catalyst, foils, etc., combined with
the fact that virtually any carbon-containing organic compound decomposes to CO2
as an end product complicates the interpretation of experimental results and eluci-
dation of the mechanism. By using isotopically labeled gases, true photocatalysis
can be distinguished from reactions where the carbon source originates from
adsorbed gasses or carbon-containing precursors.
The photocatalyst must be abundant, cheap, and chemically stable. These qual-
ities are (among other transition metal oxides) found in TiO2 possibly modified with
dopants. But catalyst preparation is an issue that receives limited attention in
connection with energy usage. Most catalysts require annealing at high tempera-
tures for several hours to form an active crystal phase – an energy costly process
that has to be repaid by catalytic action or carried out by using renewable energy,
e.g., concentrated sunlight in solar ovens. The prize of manufacturing catalysts on a
very large scale is also challenging, especially if doping with expensive metals such
as platinum is required. To enhance the reactivity extension of the absorption
(of the catalytic species) into the visible region of the electromagnetic spectrum
220 J. Jensen and F.C. Krebs

seems like a viable road forward if catalysts are to achieve conversion factors that
outdo the manufacture of the catalytic species themselves.
Dispersion of photocatalytic species onto large areas is likewise an issue that
needs addressing. We have, in this chapter, outlined the possibilities and difficulties
in using flexible substrates as a solution. Flexible substrates can be coated with
photocatalyst at a considerable rate and show promise to be compatible with
membranes for carbon dioxide extraction. Roll-to-roll (R2R) coating is a
low-tech method, which only requires moderate investments in equipment.
Construction of huge photocatalytic reactors is yet another challenge. Most
reports are on laboratory size reactors, but this is very far from what is needed as
the conditions under which tests are made. The challenge in going from the
laboratory to the industrial scale seems to be huge, but it is definitely one that
must be addressed.

References

1. Mikkelsen M, Jørgensen M, Krebs FC (2010) The teraton challenge. A review of fixation and
transformation of carbon dioxide. Energy Environ Sci 3:43–81
2. Martin LR (1980) Use of solar energy to reduce carbon dioxide. Sol Energy 24:271–277
3. Traynor AJ, Jensen RJ (2002) Direct solar reduction of CO2 to fuel: first prototype results. Ind
Eng Chem Res 41:1935–1939
4. Carp O, Huisman CL, Reller A (2004) Photoinduced reactivity of titanium dioxide. Prog Solid
State Chem 32:33–177
5. Varghese OK, Grimes CA (2008) Appropriate strategies for determining the photoconversion
efficiency of water photoelectrolysis cells: a review with examples using titania nanotube array
photoanodes. Sol Energy Mater Sol Cells 92:374–384
6. Lo C, Hung C, Yuan C, Wu J (2007) Photoreduction of carbon dioxide with H2 and H2O over
TiO2 and ZrO2 in a circulated photocatalytic reactor. Sol Energy Mater Sol Cells
91:1765–1774
7. Roy SC, Varghese OK, Paulose M, Grimes CA (2010) Toward solar fuels: photocatalytic
conversion of carbon dioxide to hydrocarbons. ACS Nano 4:1259–1278
8. Nguyen T, Wu JCS (2008) Photoreduction of CO2 to fuels under sunlight using optical-fiber
reactor. Sol Energy Mater Sol Cells 92:864–872
9. Linsebigler AL, Lu GQ, Yates JT (1995) Photocatalysis on TiO2 surfaces – principles,
mechanisms, and selected results. Chem Rev 95:735–758
10. Jensen J, Mikkelsen M, Krebs FC (2011) Flexible substrates as basis for photocatalytic
reduction of carbon dioxide. Sol Energy Mater Sol Cells 95:2949–2958
11. Deng L, Hägg M (2010) Swelling behavior and gas permeation performance of PVAm/PVA
blend FSC membrane. J Membr Sci 363:295–301
12. Nishimura A, Komatsu N, Mitsui G, Hirota M, Hu E (2009) CO2 reforming into fuel using
TiO2 photocatalyst and gas separation membrane. Catal Today 148:341–349
13. Derichter RK, Ming T, Caillol S (2013) Fighting global warming by photocatalytic reduction
of CO2 using giant photocatalytic reactors. Renew Sustain Energy Rev 19:82–106
14. Jin C, Christensen PA, Egerton TA, Lawson EJ, White JR (2006) Rapid measurement of
polymer photo-degradation by FTIR spectrometry of evolved carbon dioxide. Polym Degrad
Stab 91:1086–1096
15. Robinson AJ, Searle JR, Worsley DA (2004) Novel flat panel reactor for monitoring
photodegradation. Mater Sci Technol 20:1041–1048
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 221

16. Hong J, Zhang W, Ren J, Xu R (2013) Photocatalytic reduction of CO2: a brief review
on product analysis and systematic methods. Anal Methods 5:1086–1097
17. Su W, Fu X, Wei K, Zhang H, Lin H, Wang X, Li D (2001) Spectrum studies on titania
photocatalysts. Guang Pu Xue Yu Guang Pu Fen Xi 21:34
18. Ulagappan N, Frei H (2000) Mechanistic study of CO2 photoreduction in Ti silicalite molec-
ular sieve by FT-IR spectroscopy. J Phys Chem A 104:7834–7839
19. Fredin L, Nelander B, Ribbegård G (1974) On the dimerization of carbon dioxide in nitrogen
and argon matrices. J Mol Spectrosc 53:410–416
20. Mascetti J, Tranquille M (1988) Fourier transform infrared studies of atomic Ti, V, Cr, Fe, Co,
Ni, and Cu reactions with carbon dioxide in low-temperature matrices. J Phys Chem
92:2177–2184
21. Burch DE, Williams D (1962) Total absorptance of carbon monoxide and methane in the
infrared. Appl Optics 1:587–594
22. Falk M, Whalley E (1961) Infrared spectra of methanol and deuterated methanols in gas,
liquid, and solid phases. J Chem Phys 34:1554–1568
23. Nelander B (1980) Infrared spectrum of the water formaldehyde complex in solid argon and
solid nitrogen. J Chem Phys 72:77–84
24. Stein SE (2013) In: Linstrom J, Mallard WG (eds) NIST chemistry WebBook, NIST standard
reference database 69. NIST, Gaithersburg, 20899
25. Yang C, Yu Y, Van Der Linden B, Wu JCS, Mul G (2010) Artificial photosynthesis over
crystalline TiO2-based catalysts: fact or fiction? J Am Chem Soc 132:8398–8406
26. Lin W, Han H, Frei H (2004) CO2 splitting by H2O to CO and O2 under UV light in TiMCM-41
silicate sieve. J Phys Chem B 108:18269–18273
27. Kobayashi KLI, Shiraki Y, Katayama Y (1978) Study of adsorption phenomena on aluminium
by interatomic Auger transition spectroscopy. I. Initial oxidation of Al. Surf Sci 77:449–457
28. Parkyns ND (1967) Surface properties of metal oxides. Part I. Infrared studies of the adsorp-
tion and oxidation of carbon monoxide on alumina. J Chem Soc A Inorg Phys Theor
Chem:1910–1913. doi:10.1039/J19670001910
29. Mao C, Vannice MAS (1994) High surface area a-alumina. I. Adsorption properties and heats
of adsorption of carbon monoxide, carbon dioxide, and ethylene. Appl Catal A Gen
111:151–173
30. Shiraki Y, Kobayashi KLI, Katayama Y (1978) Study of adsorption phenomena on aluminium
by interatomic Auger transition spectroscopy. II. CO adsorption onto Al. Surf Sci 77:458–464
31. Chen C, Ahn W (2011) CO2 capture using mesoporous alumina prepared by a sol-gel process.
Chem Eng J 166:646–651
32. Morterra C, Magnacca G (1996) A case study: surface chemistry and surface structure of
catalytic aluminas, as studied by vibrational spectroscopy of adsorbed species. Catal Today
27:497–532
33. Zhou Z, Zeng T, Cheng Z, Yuan W (2011) Preparation and characterization of titania-alumina
mixed oxides with hierarchically macro-/mesoporous structures. Ind Eng Chem Res
50:883–890
34. Norrman K, Ghanbari-Siahkali A, Larsen NB (2005) Studies of spin-coated polymer films.
Annu Rep Progr Chem Sect C 101:174–201
35. Krebs FC (2009) Fabrication and processing of polymer solar cells: a review of printing and
coating techniques. Sol Energy Mater Sol Cells 93:394–412
36. Dam HF, Krebs FC (2012) Simple roll coater with variable coating and temperature control for
printed polymer solar cells. Sol Energy Mater Sol Cells 97:191–196
37. Cozzoli PD, Kornowski A, Weller H (2003) Low-temperature synthesis of soluble and
processable organic-capped anatase TiO2 nanorods. J Am Chem Soc 125:14539–14548
38. Willis AL, Turro NJ, O’Brien S (2005) Spectroscopic characterization of the surface of iron
oxide nanocrystals. Chem Mater 17:5970–5975
222 J. Jensen and F.C. Krebs

39. Satapathy S, Gupta PK, Srivastava H, Srivastava AK, Wadhawan VK, Varma KBR, Sathe VG
(2007) Effect of capping ligands on the synthesis and on the physical properties of the
nanoparticles of LiTaO3. J Cryst Growth 307:185–191
40. Liu GQ, Jin ZG, Liu XX, Wang T, Liu ZF (2007) Anatase TiO2 porous thin films prepared by
sol-gel method using CTAB surfactant. J Sol Gel Sci Technol 41:49–55
41. Kehres J, Andreasen JW, Krebs FC, Molenbroek AM, Chorkendorff I, Vegge T (2010)
Combined in situ small- and wide-angle X-ray scattering studies of TiO2 nanoparticle
annealing to 1023K. J Appl Crystallogr 43:1400
42. Huo Q, Margolese DI, Ciesla U, Feng P, Gier TE, Sieger P, Leon R, Petroff PM, Schüth F,
Stucky GD (1994) Generalized synthesis of periodic surfactant/inorganic composite materials.
Nature 368:317–321
43. Li Z, Zhu Y (2003) Study on the surface-modification of TiO2 nanoparticles. Appl Surf Sci
61:1484–1487
44. Shi X, Rosa R, Lazzeri A (2010) On the coating of precipitated calcium carbonate with stearic
acid in aqueous medium. Langmuir 26:8474–8482
45. Parkin IP, Palgrave RG (2005) Self-cleaning coatings. J Mater Chem 15:1689–1695
46. Gohin M, Allain E, Chemin N, Maurin I, Gacoin T, Boilot J (2010) Sol-gel nanoparticulate
mesoporous films with enhanced self-cleaning properties. J Photochem Photobiol A Chem
216:142–148
47. Kuai Q, Ye H, Gao Y (2010) Preparation of aluminum titanium pigments and its application in
self-cleaning coating. Fenmo Yejin Cailiao Kexue yu Gongcheng 15:283–287
48. Rathouský J, Kalousek V, Kolář M, Jirkovský J, Barták P (2011) A study into the self-cleaning
surface properties – the photocatalytic decomposition of oleic acid. Catal Today 161:202–208
49. Harwood HJ (1962) Reactions of the hydrocarbon chain of fatty acids. Chem Rev 62:99–154
50. Kang KS, Chen Y, Yoo KH, Jyoti N, Kim J (2009) Cause of slow phase transformation of TiO2
nanorods. J Phys Chem C 113:19753–19755
51. Hao B, Li Y, Wang S (2010) Synthesis and structural characterization of surface-modified
TiO2. Adv Mater Res 129–131:154–158
52. Pradhan AR, Wu JF, Jong SJ, Tsai TC, Liu SB (1997) An ex situ methodology for character-
ization of coke by TGA and 13C CP-MAS NMR spectroscopy. Appl Catal A Gen
165:489–497
53. Roonasi P, Holmgren A (2009) A Fourier transform infrared (FTIR) and thermogravimetric
analysis (TGA) study of oleate adsorbed on magnetite nano-particle surface. Appl Surf Sci
255:5891–5895
54. Gijsman P, Meijers G, Vitarelli G (1999) Comparison of the UV-degradation chemistry of
polypropylene, polyethylene, polyamide 6 and polybutylene terephthalate. Polym Degrad Stab
65:433–441
55. Gulmine JV, Janissek PR, Heise HM, Akcelrud L (2003) Degradation profile of polyethylene
after artificial accelerated weathering. Polym Degrad Stab 79:385–397
56. Hamid SH, Amin MB (1995) Lifetime prediction of polymers. J Appl Polym Sci
55:1385–1394
57. Day M, Wiles DM (1972) Photochemical degradation of poly(ethylene terephthalate). II.
Effect of wavelength and environment on the decomposition process. J Appl Polym Sci
16:191–202
58. Day M, Wiles DM (1971) Photochemical decomposition mechanism of poly(ethylene tere-
phthalate). J Pol Sci B 9:665–669
59. Kockott D (1989) Natural and artificial weathering of polymers. Polym Degrad Stab
25:181–208
60. Day M, Wiles DM (1972) Photochemical degradation of poly(ethylene terephthalate). III.
Determination of decomposition products and reaction mechanism. J Appl Polym Sci
16:175–189
61. Philippart J, Posada F, Gardette J (1995) Mass spectroscopy analysis of volatile photoproducts
in photooxidation of polypropylene. Polym Degrad Stab 49:285–290
8 Fixation of Carbon Dioxide Using Molecular Reactions on Flexible Substrates 223

62. Fernando SS, Christensen PA, Egerton TA, Eveson R, Martins-Franchetti SM, Voisin D,
White JR (2009) Carbon dioxide formation during initial stages of photodegradation of poly
(ethyleneterephthalate) (PET) films. Mater Sci Technol 25:549–555
63. Grossetete T, Rivaton A, Gardette JL, Hoyle CE, Ziemer M, Fagerburg DR, Clauberg H (2000)
Photochemical degradation of poly(ethylene terephthalate)-modified copolymer. Polymer
41:3541–3554
64. Fechine GJM, Souto-Maior RM, Rabello MS (2007) Photodegradation of multilayer films
based on PET copolymers. J Appl Polym Sci 104:51–57
65. Hurley CR, Leggett GJ (2009) Quantitative investigation of the photodegradation of polyeth-
ylene terephthalate film by friction force microscopy, contact-angle goniometry, and X-ray
photoelectron spectroscopy. ACS Appl Mater Interfaces 1:1688–1697
66. Chew CH, Gan LM, Scott G (1977) Mechanism of the photo-oxidation of polyethylene. Eur
Polym J 13:361–364
67. Gijsman P, Dozeman A (1996) Comparison of the UV-degradation chemistry of unstabilized
and HALS-stabilized polyethylene and polypropylene. Polym Degrad Stab 53:45–50
68. Costa L, Luda MP, Trossarelli L (1997) Ultra high molecular weight polyethylene –
II. Thermal- and photo-oxidation. Polym Degrad Stab 58:41–54
69. Lacoste J, Deslandes Y, Black P, Carlsson DJ (1995) Surface and bulk analyses of the
oxidation of polyolefins. Polym Degrad Stab 49:21–28
70. Arnaud R, Moisan J, Lemaire J (1984) Primary hydroperoxidation in low-density polyethyl-
ene. Macromolecules 17:332–336
71. Rivaton A (1993) Photochemistry of poly(butyleneterephthalate): 2-Identification of the
IR-absorbing photooxidation products. Polym Degrad Stab 41:297–310
72. Rivaton A (1993) Photochemistry of poly(butyleneterephthalate): 1-Identification of the
IR-absorbing photolysis products. Polym Degrad Stab 41:283–296
73. Chong MN, Jin B, Chow CWK, Saint C (2010) Recent developments in photocatalytic water
treatment technology: a review. Water Res 44:2997–3027
74. Fujishima A, Zhang X, Tryk DA (2008) TiO2 photocatalysis and related surface phenomena.
Surf Sci Rep 63:515–582
75. Hufschmidt D, Liu L, Selzer V, Bahnemann D (2004) Photocatalytic water treatment: funda-
mental knowledge required for its practical application. Water Sci Technol 49:135–140
76. Allen NS, Edge M, Corrales T, Catalina F (1998) Stabiliser interactions in the thermal and
photooxidation of titanium dioxide pigmented polypropylene films. Polym Degrad Stab
61:139–149
77. Fechine GJM, Rabello MS, Souto-Maior RM (2002) The effect of ultraviolet stabilizers on the
photodegradation of poly(ethylene terephthalate). Polym Degrad Stab 75:153–159
78. Allen NS, Khatami H, Thompson F (1992) Influence of titanium dioxide pigments on the
thermal and photochemical oxidation of low density polyethylene film. Eur Polym J
28:817–822
Chapter 9
Reductive Conversion of Carbon Dioxide
Using Various Photocatalyst Materials

Kojirou Fuku, Kohsuke Mori, and Hiromi Yamashita

9.1 Introduction

Although advances in science and technology have contributed to the development of


society and have made people’s lives increasingly comfortable, various environmen-
tal problems associated with technological advances have become global issues. In
particular, global warming by greenhouse effect gas such as carbon dioxide (CO2)
and exhaustion of fossil fuels are already grave social problems of global dimensions.
The catalytic conversion of CO2 has widely been examined as one of the valuable
approaches to solve these problems, but most of the processes require high temper-
ature and pressures. However, various photocatalytic reactions have recently
attracted considerable attention due to their milder reaction conditions (at room
temperature and atmospheric pressure) and their ability to utilize solar light as
inexhaustible energy source. Highly efficient and selective photocatalytic systems
driven under sunlight and accompanied with a large positive change in the Gibbs free
energy is of vital interest [1–8]. These reactions are considered to be an uphill
reaction and are similar to the photosynthesis by green plants which produces glucose
and oxygen from CO2. The photocatalytic conversion of CO2 emitted in large
amounts into industrially beneficial compounds would be an alternative protocol
from the viewpoints of green and sustainable chemistry by the solution of exhaustion
of energy resources and global warming. Therefore, the development of efficient

K. Fuku
Division of Materials and Manufacturing Science, Graduate School of Engineering,
Osaka University, 2-1 Yamada-oka, Suita, Osaka 565-0871, Japan
K. Mori • H. Yamashita (*)
Division of Materials and Manufacturing Science, Graduate School of Engineering,
Osaka University, 2-1 Yamada-oka, Suita, Osaka 565-0871, Japan
Elements Strategy Initiative for Catalysts Batteries ESICB, Kyoto University,
Katsura, Kyoto 615-8520, Japan
e-mail: yamashita@mat.eng.osaka-u.ac.jp

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 225
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_9,
© Springer-Verlag Berlin Heidelberg 2014
226 K. Fuku et al.

photocatalytic systems enabling reduction and/or fixation of CO2 is the most


desirable and challenging goal [2, 3].
The utilization of solar energy for the conversion of CO2 can be realized by
considering the photocatalytic reduction and/or fixation of CO2 into CO, HCOOH,
CH3OH, and CH4, etc., by the active photocatalysts such as TiO2, SrTiO3, ZnO, or
SiC semiconductors. Inoue and Fujishima et al. have first reported that HCOOH,
HCHO, and CH3OH are produced under irradiation of aqueous suspension systems
involving a variety of semiconductor powders or single crystals [9]. Although the
pioneering works on the photoreduction of CO2 on semiconductors in aqueous
suspension systems were summarized by Halmann, the efficiency was low when
H2O was used as the reductant [10]. Substantially improved yields of CH3OH and
CH4 from CO2 with H2O have been reported by us, in which photocatalytic
reactions successfully proceed in solid–gas systems at room temperature on isolated
tetrahedral Ti centers of macro- or mesoporous silicate sieves instead of dense
phase powdered TiO2 materials [2, 3, 11–22]. We focus on the characteristic
features of photocatalytic reduction and/or fixation of CO2 with H2O on various
types of active titanium oxide catalysts under heterogeneous gas–solid conditions.
This review report about our approach on the efficient reduction and/or fixation
of CO2 using various titanium oxide photocatalysts. Additionally, other novel and
unique systems enabling efficient conversion of CO2 using various oxide semi-
conductors other than titanium oxide or localized surface plasmon resonance-
enhanced photocatalysis are also introduced.

9.2 Various Titanium Oxide Photocatalysts

As shown in Fig. 9.1, the electronic properties as well as the photocatalytic activity of
titanium oxide photocatalysts dramatically change depending on their local structure,
from an extended semiconducting structure to nano-sized, extremely small
nanoparticles and molecular-sized titanium oxide species. In the case of the bulk
TiO2 materials, photochemical excitation leads to charge separation in the particle, in
which electrons (e) are promoted to the conduction band (cb) and holes (h+) are left
in the valence band (vb) upon light irradiation. The generated hole in the valence
band and the conduction band electron can react with electron donors and electron
acceptors adsorbed on the titanium oxide surface, respectively. On the contrary, the
Ti-oxide moieties, which are spatially separated from each other, can be implanted
and isolated in the silica matrixes of microporous zeolite and mesoporous silica
materials at the atomic level, and have been named as “single-site photocatalysts.”
The isolated Ti atoms can be substituted with Si atoms in the silica matrixes
(Si/Ti > 30) and coordinated tetrahedrally with oxygen atoms. Such single-site
catalysts can be simply synthesized by various anchoring techniques, such as
hydrothermal synthesis, sol–gel method, and chemical vapor deposition, and are
stable as far as the silica matrixes keep their porous structures. In the case of the
isolated and tetrahedrally coordinated Ti-oxide moiety, excitation by UV irradiation
brings about an electron transfer from the oxygen (O2) to Ti4+ ions, resulting in the
9 Reductive Conversion of Carbon Dioxide Using Various Photocatalyst Materials 227

Fig. 9.1 The electronic state change in the titanium oxide photocatalysts from semiconducting
bulk TiO2 to the isolated Ti-oxide molecular species

Fig. 9.2 Reaction scheme


for the photocatalytic
reduction of CO2 with H2O
on bulk TiO2

formation of pairs of trapped hole centers (O) and electron centers (Ti3+). Such
charge-transfer excited state, i.e., the excited electron–hole pair state which localizes
quite near to each other as compared to the electron and hole produced in semicon-
ducting materials, plays a significant role in various photocatalytic reactions.
The efficiency for the photocatalytic reduction and/or fixation of CO2 with H2O
to produce CH4 and CH3OH strongly depends upon the type of the employed
titanium oxide photocatalysts [11–22]. The primary processes on semiconducting
TiO2 photocatalysts are illustrated in Fig. 9.2. The band gap between the conduc-
tion band and the valence band becomes larger as the particle size of the semicon-
ducting TiO2 decreased, making it suitable and applicable to the reduction of CO2
[2, 3]. For extremely fine TiO2 particles less than 10 nm in diameter, the size
quantum effect and/or the effects of the surface modification in its coordination
geometry plays a significant role in the appearance of unique activity. Conse-
quently, the electrons and holes which are produced by UV-light irradiation within
the ultrafine particles of TiO2 and the highly dispersed titanium oxide species
exhibit more unique and high activities compared to those produced in large
particle TiO2 photocatalysts.
228 K. Fuku et al.

9.3 Small Particle TiO2 Photocatalysts

UV irradiation of the powdered TiO2 catalysts in the presence of a gaseous mixture


of CO2 and H2O led to the evolution of CH4 at 275 K, accompanied with the
formation of trace amounts of C2H4 and C2H6 [13, 14]. The yields of these products
increased with the UV-irradiation time, while no products were detected under dark
conditions. The CH4 yield was almost zero in the reaction of CO2 without H2O and
increased with increasing the amount of H2O. These results suggest that the
photocatalytic reduction of CO2 to produce CH4 and C2 compounds takes place
photocatalytically in the solid–gas phase systems. The formation of CH4 as
the main product has also been observed by Saladin et al., and the partially reduced
TiO2 species formed under UV irradiation is proposed as an active species [23].
The yields of CH4 formation in the photocatalytic reduction of CO2 with H2O on
several TiO2 photocatalysts with different physicochemical property are shown in
Table 9.1. The photocatalytic activity, based on the CH4 yields, was found to
depend on the type of TiO2 catalyst, in the order of JRC-TIO-4 > -5 > -2 > -3 [14].
This tendency corresponds with that in the photocatalytic hydrogenation of
methyl acetylene with H2O, proving that the reduction of CO2 with H2O
undoubtedly occurs photocatalytically activated powdered TiO2 catalyst [24, 25]. It
is likely that the anatase-type TiO2 possessing a large band gap as well as numerous
surface –OH groups is preferable for efficient photocatalytic reactions. The band gap
increase is accompanied by a shift in the conduction band edge to higher energy levels.
This shift causes the reductive potential to shift to more negative values, which
ultimately causes a great enhancement in the photocatalytic activity. The surface –OH
groups and/or physisorbed H2O also play a significant role in photocatalytic
reactions via the formation of OH radicals and H radicals.
Preliminary studies by ESR measurement provide an insight into the reaction
pathway. The ESR signals obtained under UV irradiation of the anatase-type TiO2
catalyst in the presence of CO2 and H2O at 77 K can be ascribed to the characteristic
photogenerated Ti3+ ions (g⊥ ¼ 1.9723 and gk ¼ 1.9628) and H radicals (with
490 G splitting), as well as CH3 radicals having a hyperfine splitting (Hα ¼ 19.2 G,
g ¼ 2.002) [14]. The signal intensity of CH3 radicals decreased with increasing the
amount of H2O, indicating that CH3 radicals are the intermediate species and react
with H radicals that are formed by the reduction of protons (H+) originated from
H2O adsorbed on the catalyst.

Table 9.1 Textural property and photocatalytic activity of TiO2 catalysts


Band Reduction Hydrogenation of
Crystal SBET CO2 ads. Relative gap of CO2b methyl acetylene
Catalyst structurea (m2/g) (μmol/g) –OH conc. (eV) (μmol/h·g) (μmol/h·g)
JRC-TIO-2 A 16 1 1 3.47 0.03 0.20
JRC-TIO-3 R 51 17 1.6 3.32 0.02 0.12
JRC-TIO-4 A 49 10 3.0 3.50 0.27 8.33
JRC-TIO-5 R 3 0.4 3.1 3.09 0.04 0.45
a
A, anatase; R, rutile
b
CH4 yield in the reaction of CO2 (0.12 mmol) and H2O (0.37 mmol) for 6 h
9 Reductive Conversion of Carbon Dioxide Using Various Photocatalyst Materials 229

9.4 Metal-Loaded TiO2 Photocatalysts

The effect of metal loading on the photocatalytic reduction of CO2 with H2O on
TiO2 catalysts was investigated [14]. In the case of the Cu/TiO2 photocatalyst
(0.3–1.0 wt%), CH4 yield was suppressed, but a new formation of CH3OH could
be observed. Characterization by XPS reveals that the main species of copper in the
catalyst is 1+ oxidation state. It has been also reported that Cu+ catalysts play a
significant role in the photoelectrochemical production of CH3OH from CO2 and
H2O system [26]. Following the photocatalytic reduction, the Cu/TiO2 catalysts
exhibited a new peak at 299 eV in the C(1) XPS spectra, suggesting that carboxylate
groups which accumulated on the catalyst may be the primary intermediate species
for this reaction.
Further detailed study was performed by Gunlazuardi et al. by systematically
varying the Cu loading on TiO2 [27]. The CH3OH yield increased with Cu loading,
and the highest yield can be achieved by 3 % Cu/TiO2, which was three times
higher than the original TiO2. Cu can serve as an electron trapper and prohibits the
recombination of electron and hole, significantly increasing photoefficiency. How-
ever, catalysts with more than 3 wt% Cu loading do not further increase the CH3OH
yield due to its shading effects, consequently reducing the photo-exciting capacity
of TiO2. The activation energy for 3 % Cu/TiO2 and the original TiO2 was
determined to be 12 and 26 kJ/mol, respectively. The apparent lower activation
energy of 3 % Cu/TiO2 catalyst indicates that the Cu also acts as active species to
provide CH3OH and enhances the photoefficiency of TiO2 photocatalysts.
In the case of the Pt/TiO2, the yield of CH4 increased remarkably when the
amount of Pt was increased (0.1–1.0 wt%), but the addition of excess Pt was
undesirable for an efficient reaction [13]. Regarding the reaction intermediates,
Solymosi et al. have observed the formation of CO2 species in bent form under
UV irradiation of the Rh/TiO2 in the presence of CO2 using FT-IR [28]. The
electron transfer from the irradiated catalyst to the adsorbed CO2 takes place,
resulting in the formation of a CO2 anion as the key step in the photochemical
reduction of CO2 using metal-loaded TiO2 semiconductor.

9.5 TiO2 Single Crystals

With a well-defined catalyst surface such as a single crystal, detailed information on


the reaction mechanism can be obtained at the molecular level [2, 15]. Therefore, the
photocatalytic reduction of CO2 with H2O on rutile-type TiO2 (100) and TiO2 (110)
single crystal surfaces has been performed [15]. As shown in Table 9.2, the efficiency
and selectivity of the photocatalytic reactions strongly depend on the type of TiO2
single crystal surface. UV irradiation of the TiO2 (100) single crystal catalyst in the
presence of a mixture of CO2 and H2O led to the evolution of both CH4 and CH3OH at
275 K, whereas only CH3OH was detected with the TiO2 (110) single crystal catalyst.
230 K. Fuku et al.

Table 9.2 Yields of the formation of CH4 and CH3OH in the photocatalytic reduction of CO2
(124 μmol g1) with H2O (372 μmol g1) at 275 K
Single crystal of TiO2 Yield of CH3OH (μmol/h∙g-cat) Yield of CH4 (μmol/h·g-cat)
(100) 2.4 3.5
(110) 0.8 0

It is likely that the photogenerated electrons localizing on the surface sites of the
excited TiO2 play a significant role in the photoreduction of CO2 molecules into
intermediate carbon species [2, 15]. The surface Ti atoms may act as a reductive site.
According to the surface geometric models for TiO2 (100) and TiO2 (110), the atomic
ratio (Ti/O) of the top-surface Ti and O atoms, which have geometric spaces large
enough to have direct contact with CO2 and H2O molecules, is higher on TiO2 (100)
than on TiO2 (110) surface. In the excited state, the surface with a higher Ti/O surface
ratio, i.e., TiO2 (100), exhibits a more reductive tendency than TiO2 (110). Such a
reductive surface allows a more facile reduction of CO2 molecules especially for the
formation of CH4. In the HREELS spectrum of a clean TiO2(100) single crystal
surface after UV irradiation in the presence of CO2 and H2O, two peaks due to the
C–H stretching vibration of the CHx species and the O–H stretching of the surface
hydroxyl groups at around 2,920 and 3,630 cm1, respectively. On the other hand,
only a weak peak assigned to the O–H stretching vibration was observed without UV
irradiation, suggesting that UV-light irradiation is indispensable for attaining CO2
reduction and the formation of the active H and CHx species.

9.6 Ti-Oxide Anchored on Zeolite (Ion Exchange)

The photocatalyst systems incorporated within the zeolite cavities and frameworks
have been proven to be effective for various reactions [29–32]. The Ti-oxide
anchored onto zeolite, Ti-oxide/Y-zeolite (1.1 wt% as TiO2), was prepared by ion
exchange with an aqueous titanium ammonium oxalate solution using Y-zeolite
(SiO2/Al2O3 ¼ 5.5) (ex-Ti-oxide/Y-zeolite) [16–18].
Figure 9.3 shows the Ti K-edge XANES and the Fourier transformation of EXAFS
(FT-EXAFS) spectra of the Ti-oxide/Y-zeolites. The XANES spectra of the bulk TiO2
powder give rise to several well-defined pre-edge peaks attributable to the titanium in
a symmetric octahedral environment [33]. The ex-Ti-oxide/Y-zeolite exhibits an
intense single pre-edge peak at 4,967 eV, suggesting that the Ti-oxide species exist
in a tetrahedral coordination [1–3]. On the other hand, the imp-Ti-oxide/Y-zeolite
prepared by the impregnation exhibits three characteristic weak pre-edge peaks
attributed to crystalline TiO2. The FT-EXAFS spectra of the ex-Ti-oxide/Y-zeolite
exhibit only peak at around 1.6 Å assigned to the neighboring oxygen atoms (Ti–O)
indicating the presence of an isolated Ti-oxide species. These findings indicate that
highly dispersed isolated tetrahedral Ti-oxide species are formed on the ex-Ti-oxide/
Y-zeolite. On the other hand, the imp-Ti-oxide/Y-zeolite exhibits an intense peak
9 Reductive Conversion of Carbon Dioxide Using Various Photocatalyst Materials 231

Fig. 9.3 Ti K-edge XANES and FT-EXAFS spectra of anatase TiO2 powder (a), imp-Ti-oxide/
Y-zeolite (10.0 wt% as TiO2) (b), and the ex-Ti-oxide/Y-zeolite (c)

assigned to the neighboring titanium atoms (Ti–O–Ti) suggestive of the aggregation


of the Ti-oxide species.
The photoluminescence spectra of the ex-Ti-oxide/Y-zeolite at 77 K are shown in
Fig. 9.4. Excitation by light at around 250–280 nm brought about an electron transfer
from the oxygen to titanium ion, resulting in the formation of pairs of the trapped hole
center (O) and an electron center (Ti3+) [1–3]. The observed photoluminescence is
attributed to the radiative decay process from the charge-transfer excited state of the
Ti-oxide moieties in a tetrahedral coordination geometry, (Ti3+—O)*, to their ground
state [34, 35]. The addition of H2O or CO2 molecules onto the anchored Ti-oxide
species leads to the efficient quenching of the photoluminescence. Such an efficient
quenching suggests not only that tetrahedrally coordinated Ti-oxide species locate at
positions accessible to the added CO2 or H2O but also that CO2 or H2O interacts and/or
reacts with the Ti-oxide species in both its ground and excited states. Because the
addition of CO2 led to a less effective quenching than that with the addition of H2O, the
interaction of the emitting sites with CO2 was weaker than that with H2O.
UV irradiation of powdered TiO2 and Ti-oxide/Y-zeolite catalysts in the presence
of a mixture of CO2 and H2O led to the evolution of CH4 and CH3OH at 328 K, as
well as trace amounts of CO, C2H4, and C2H6 [16–18]. The specific photocatalytic
activities for the formation of CH4 and CH3OH are shown in Fig. 9.5. The ex-Ti-
oxide/Y-zeolite exhibits a high activity and a high selectivity for the formation of
CH3OH, while the formation of CH4 was predominated on bulk TiO2 as well as on
232 K. Fuku et al.

Fig. 9.4 Photoluminescence spectrum of the ex-Ti-oxide/Y-zeolite catalyst (a) and the effects of
the addition of CO2 and H2O. Measured at 77 K, excitation at 290 nm, and emission monitored at
490 nm

the imp-Ti-oxide/Y-zeolite. The deposition of Pt improved the photocatalytic


activity, but the CH3OH selectivity significantly decreased. These findings clearly
suggest that the tetrahedrally coordinated Ti-oxide species act as active
photocatalysts for the reduction of CO2 with H2O exhibiting a high selectivity toward
CH3OH.
UV irradiation of the anchored Ti-oxide catalyst in the presence of CO2 and H2O
at 77 K led to the appearance of ESR signals due to the Ti3+ ions, H atoms, and carbon
radicals [2, 3]. After the disappearance of these ESR signals in this system at around
275 K, the formation of CH4 and CH3OH was observed. From these results, the
reaction mechanism in the photocatalytic reduction of CO2 with H2O on the highly
dispersed Ti-oxide catalyst can be proposed as shown in Scheme 9.1: CO2 and H2O
molecules interact with the excited state of the photoinduced (Ti3+—O)* species
and the reduction of CO2 and the decomposition of H2O proceed competitively.
Furthermore, H atoms and OH• radicals are formed from H2O and such radicals react
with the carbon species formed from CO2 to produce CH4 and CH3OH.
9 Reductive Conversion of Carbon Dioxide Using Various Photocatalyst Materials 233

Fig. 9.5 The product distribution of the photocatalytic reduction of CO2 with H2O on anatase
TiO2 powder, the imp-Ti-oxide/Y-zeolite (10 wt% as TiO2), the imp-Ti-oxide/Y-zeolite (1 wt% as
TiO2), the ex-Ti-oxide/Y-zeolite (1 wt% as TiO2), and the Pt-loaded ex-Ti-oxide/Y-zeolite
catalysts

Scheme 9.1 Schematic illustration of the photocatalytic reduction of CO2 with H2O on the
anchored titanium oxide species
234 K. Fuku et al.

9.7 Ti-Containing Zeolite and Mesoporous


Molecular Sieves

The Ti-oxide species prepared within the ordered silica frameworks have revealed a
unique local structure as well as a high selectivity in the oxidation of organic
substances with hydrogen peroxide [34–36]. Hydrothermally synthesized
Ti-containing zeolites (TS-1, Ti-Beta) and mesoporous molecular sieves (Ti-MCM,
Ti-HMS, Ti-FSM) have been subjected to the photocatalytic reduction CO2 with H2O
[18–22, 36–38].
In situ photoluminescence, ESR, UV–VIS and XAFS investigations indicated
that the Ti-oxide species in the Ti-mesoporous molecular sieves (Ti-MCM-41 and
Ti-MCM-48) and in the TS-1 zeolite are highly dispersed within the zeolite
framework and exist in a tetrahedral coordination. Upon excitation with UV light
at around 250–280 nm, these catalysts exhibit photoluminescence spectra at around
480 nm. The addition of CO2 or H2O onto these catalysts results in a significant
quenching of the photoluminescence, suggesting the excellent accessibility of the
CO2 and H2O to the Ti-oxide species [18–21].
UV irradiation of the Ti-mesoporous molecular sieves and the TS-1 zeolite in the
presence of CO2 and H2O also led to the formation of CH3OH and CH4 as the main
products [18–21]. The yields of CH3OH and CH4 per unit weight of the Ti-based
catalysts are shown in Fig. 9.6. It can be seen that Ti-MCM-48 exhibits much higher
activity than either TS-1 or Ti-MCM-41. The higher activity and selectivity for the
formation of CH3OH attained with the Ti-MCM-48 than that with the other catalysts
may be due to the combined contribution of the high dispersion state of the Ti-oxide
species and the large pore size with a three-dimensional channel structure: TS-1 has a
smaller pore size (ca. 5.7 Å) and a three-dimensional channel structure; Ti-MCM-41
has a large pore size (>20 Å) but one-dimensional channel structure; and
Ti-MCM-48 has both a large pore size (>20 Å) and three-dimensional channels.
These results strongly suggest that mesoporous molecular sieves containing highly
dispersed Ti-oxide species are promising candidates as effective photocatalysts.
The effect of Pt loading on the photocatalytic activity of Ti-containing
mesoporous silica has also been investigated, and the changes in the yields of
CH4 and CH3OH formation are shown in Fig. 9.6. The deposition of Pt onto the
Ti-containing zeolites is effective for promoting the photocatalytic activity, but the
formation of only CH4 is promoted [18].
Ti-containing zeolite, Ti-Beta, has attracted much attention because of their
large-pore structure compared to the Y-zeolite [36–38]. The other characteristic
feature is that the H2O affinity of Ti-Beta zeolites can be changed significantly
depending on the preparation methods, and their hydrophobic–hydrophilic proper-
ties can modify the catalytic performances [36–38]. As shown in Fig. 9.7, the
photocatalytic reduction of CO2 with H2O was found to proceed in the gas phase
at 323 K with different activity and selectivity on hydrophilic Ti-Beta(OH) and
hydrophobic Ti-Beta(F) zeolites prepared in the OH and F media, respectively.
9 Reductive Conversion of Carbon Dioxide Using Various Photocatalyst Materials 235

Fig. 9.6 The product


distribution of the
photocatalytic reduction of
CO2 with H2O on TiO2
powder, TS-1, Ti-MCM-41,
Ti-MCM-48, and the
Pt-loaded Ti-MCM-48
catalysts

Fig. 9.7 The product


distribution of the
photocatalytic reduction
of CO2 with H2O on
Ti-Beta(F), Ti-Beta(OH),
and TiO2 powder (P-25)
as the reference catalyst

The higher activity for the formation of CH4 observed with Ti-Beta(OH) and the
higher selectivity for the formation of CH3OH observed with the Ti-Beta(F) may be
attributed to the different abilities of zeolite pores on the H2O affinity. These results
suggest that the affinity of the H2O molecules to adsorb on the zeolite is one of
important factors for the selectivity in the photocatalytic reduction of CO2 and H2O.
For the reduction of CO2 with H2O on the Ti-containing zeolite and mesoporous
molecular sieves, a similar reaction pathway with Ti-oxide anchored on zeolite is
proposed by considering ESR spectroscopy at 77 K, in which CO2 reduction and
236 K. Fuku et al.

H2O splitting proceed competitively at the LMCT-excited Ti—O centers: CO2 is


reduced to CO and subsequently to C radicals, while H2O photodecomposes to H
and OH radicals. Reaction of H and OH radicals with carbon species is thought to
yield CH3OH and CH4. Frei and co-workers carried out further mechanistic inves-
tigations by monitoring the visible light-induced reaction of 13CO2 and H2O gas
mixtures in the framework Ti-MCM-41 molecular sieve with in situ FT-IR spec-
troscopy at room temperature [39]. 12CO gas along with 13CO were observed as the
products by infrared, and the growth of 13CO depended linearly on the photolysis
laser power. This points out the presence of small amounts of carbonaceous
residues on the high-surface area mesoporous silicates. H2O was also confirmed
as the stoichiometric electron donor. These results suggest that CO is a single-
photon, 2-electron-transfer product of CO2 at framework Ti centers with H2O
acting as a direct electron donor. By considering this, the mechanism is proposed
as follows: Excitation of the framework Ti4+ centers leads to photoinduced
(Ti3+—O)* species. Electron transfer from transient Ti3+ to CO2 splits the mole-
cule into CO and O. The latter is spontaneously protonated by a Si–OH group or
H+ cogenerated upon H2O oxidation to yield a surface OH radical. Another surface
OH radical is formed as a result of the simultaneous H2O oxidation by the
framework oxygen hole. The OH radicals either combine to yield H2O2 or
dismutate to give O2 and H2O.

9.8 Ti-Oxide Anchored on Porous Silica Glass (CVD)

The Ti-oxide anchored onto porous silica glass (PVG) plate was prepared using a
facile reaction of TiCl4 with the surface OH groups on the transparent porous Vycor
glass (Corning Code 7930) in the gas phase at 453–473 K, followed by treatment
with H2O vapor to hydrolyze the anchored compound [11–13]. UV irradiation of
the anchored Ti-oxide catalysts in the presence of a mixture of CO2 and H2O led to
the evolution of CH4, CH3OH, and CO at 323 K. The total yield was larger under
UV irradiation at 323 K than at 275 K. The efficiency of the photocatalytic reaction
strongly depends on the ratio of H2O/CO2, and its activity increases with increasing
the H2O/CO2 ratio; however, an excess amount of H2O suppresses the reaction
rates. Figure 9.8 shows the effect of the number of anchored Ti–O layers on the
absorption edge in the UV–VIS spectra of the catalysts and the efficiency of the
photocatalytic reactions as well as the relative yields of the photoluminescence. It
was proven that only catalysts with highly dispersed monolayer Ti-oxide exhibit
high photocatalytic activity and photoluminescence at around 480 nm. Only the
tetrahedrally coordinated Ti-oxide species exhibit photoluminescence emission
upon excitation at around 250–280 nm. These findings also clearly suggest that
the tetrahedrally coordinated Ti-oxide species function as active photocatalysts for
the reduction of CO2 with H2O.
9 Reductive Conversion of Carbon Dioxide Using Various Photocatalyst Materials 237

Fig. 9.8 The effects of the number of the Ti–O layers of the anchored titanium oxide catalysts on
the absorption edge of the catalysts (a), the reaction yields (b), and the relative yields of the
photoluminescence (c)

9.9 Ti-Containing Porous Silica Thin Film

In the syntheses of zeolites and mesoporous molecular sieves, the morphology of


these materials was mainly in powdered form. However, powdered material sys-
tems are difficult to handle for practical use as photocatalysts. Therefore, the
synthesis of transparent porous silica thin films is a subject of current interest.
Such transparent porous silica thin films have a larger surface area, in contrast to
metal oxide thin films on a quartz substrate, and can realize the efficient absorption
of light, showing that they have potential for use as effective photocatalysts.
Self-standing Ti-containing porous silica (PS) thin films with different pore
structures and Ti contents were synthesized by a solvent evaporation method.
Ti-PS(h, 25) and Ti-PS(c, 50) mean hexagonal and cubic structure with a Si/Ti
ratio of 25 and 50, respectively [40, 41]. The synthesized thin films are colorless
and completely transparent. Mesoporous structure was confirmed by a characteris-
tic diffraction peak at around 2–3 associated with the d100 spacing.
UV irradiation of these Ti-containing porous silica thin films in the presence
of CO2 and H2O at 323 K led to the formation of CH4 and CH3OH as well as
CO and O2 as minor products. Ti-PS(h, 50) thin film was proven to be the most
effective photocatalyst with a high quantum yield of 0.28 %, which was larger than
the Ti-MCM-41 powder catalyst even with the same pore structure (Fig. 9.9). This
quantum yields obtained with transparent Ti-PS (50) thin film photocatalyst has
been improved in contrast to the value obtained with the titanium oxide anchored on
transparent porous silica glass (0.02 %). In powdered form, the effect of the
scattering of light on the particle surface would be large so that effective light
238 K. Fuku et al.

Fig. 9.9 The product


distribution of the
photocatalytic reduction of
CO2 with H2O and amount
of surface OH groups on
Ti-PS(h, 25), Ti-PS(h, 50),
powdered Ti-MCM-41, and
Ti-PS(c, 50)

absorption and measurement may not be realized. Thus, the high photocatalytic
activity of such thin films can be attributed to the efficient absorption of UV light
due to its high transparency. From FT-IR investigations, it was found that these
Ti-containing porous silica thin films had different concentrations of surface OH
groups and showed different adsorption properties for the H2O molecules toward
the catalyst surface: photocatalysts having small amounts of surface OH groups
showed high selectivity for the formation of CH3OH.

9.10 Ti/Si Binary Oxide (Sol–Gel)

Ti/Si binary oxides involving different Ti contents were prepared by the sol–gel
method using mixtures of tetraethylorthosilicate and titanium isopropoxide. Ti/Si
binary oxides with only a small Ti contents exhibited the photoluminescence emis-
sion at around 480 nm upon excitation at around 280 nm [22, 42]. UV irradiation of
the Ti/Si binary oxide catalysts in the presence of a gaseous mixture of CO2 and H2O
led to the formation of CH4 and CH3OH as the main products. A parallel relationship
between the specific photocatalytic activities of the titanium oxide species and the
photoluminescence yields of the Ti/Si binary oxides was observed. This phenomenon
clearly indicates that the appearance of high photocatalytic activity for the binary
oxides is closely associated with the formation of the charge-transfer excited complex
due to the highly dispersed tetrahedral Ti-oxide species.
The XAFS, ESR, and photoluminescence investigations of the Ti/Si binary
oxide indicated that these catalysts prepared by the sol–gel method can keep
tetrahedral coordination geometry of Ti-oxide species until the TiO2 content
approaches up to approximately 20 wt%. Consequently, the Ti/Si binary oxides
having a high Ti content can be successfully utilized as active photocatalysts for the
efficient reduction of CO2 with H2O in the gas–solid system.
9 Reductive Conversion of Carbon Dioxide Using Various Photocatalyst Materials 239

9.11 Visible Light-Sensitive TiO2-Based Catalysts

The development of visible light-sensitive photocatalysts was intensively pursued


since the pure TiO2 and Ti-oxide species only shows fascinating photocatalytic
activities under UV-light irradiation, which limit the utilization of a small UV
fraction of natural solar light. One of the most promising strategies to use higher
wavelength of the solar spectrum is the dye sensitization of TiO2. The organome-
tallic dye, especially (2,20 -bipyridine) ruthenium (II) chloride hexahydrate, attached
TiO2 absorb sunlight and injects electrons into the conduction band of TiO2, which
ultimately promote the CO2 reduction to produce CH4 [43]. It was also reported that
the perylene diimide derivatives have shown high light-harvesting capacity similar
to the Ru complex dye.
Nguyen et al. found substantial improvements in the photoactivity of metal-doped
TiO2–SiO2 mixed oxide-based photocatalysts toward CO2 photoreduction in the
presence of H2O and concentrated sunlight in an optical fiber photoreactor (OFPR)
[44]. Fe atoms are introduced into the TiO2–SiO2 lattice during sol–gel process,
resulting in the full visible light absorption as well as the effect on product selectivity
of the obtained catalyst. Cu–Fe/TiO2 catalyst mainly produced ethylene with the
quantum yield of 0.024 %, whereas Cu–Fe/TiO2–SiO2 catalyst exhibited favorable
methane production with the quantum yield of 0.05 %. The overall energy efficiency
is found to be higher on Cu–Fe/TiO2–SiO2 (0.018 %) than on its Cu–Fe/TiO2
(0.016 %). The superior photoactivity of Cu–Fe/TiO2–SiO2 catalyst under natural
sunlight could be ascribed to the efficient charge-transfer mechanism between TiO2
and Cu as well as Fe as co-dopants and its full absorption of visible light.

9.12 Various Metal Oxide Photocatalysts

The reviews concerning the photocatalytic reduction of CO2 using various metallic
oxides other than TiO2 have been previously summarized by Navalon et al. [45] In
general, ZnO and NiO are widely known as versatile oxide semiconductors except
for TiO2. Because these ZnO and NiO semiconductors possess negatively high
conduction band potential compared to the TiO2 (TiO2, 0.29 V; ZnO, 0.31 V;
NiO, 0.50 V at pH 7) [46]. Yahaya et al. focused on these conduction band
potential for efficient CO2 reduction, and the photocatalytic reduction of CO2 to
CH3OH was performed in an aqueous suspension of these oxide semiconductors
under laser light irradiation at 355 nm [47]. The ZnO and NiO semiconductors
exhibited more efficient CH3OH production performance than the TiO2. The
amounts of produced CH3OH decreased with irradiation time when TiO2 and
ZnO were used as photocatalysts, indicating that the produced CH3OH was further
oxidized to HCHO, HCOOH, or CO2 by the high oxidation ability of these
materials, while no decrease of produced CH3OH was confirmed in the presence
of the NiO semiconductor. In addition, they discovered that more efficient CH3OH
240 K. Fuku et al.

production was achieved to convert CO2 into H2CO3, because the pH of the system
decreases by production of H2CO3 (dissolution of CO2), resulting in the shift of
conduction band edge to more positive value.
Novel oxide photocatalysts by composite compounds of various metal species
were also reported for photoreduction of CO2 [48]. Various cocatalyst nanoparticles
such as NiOx, Ru, Cu, Au, or Ag were loaded on ALa4Ti4O15 (A ¼ Ca, Sr, or Ba).
Among those synthesized, 2 wt% Ag loaded on BaLa4Ti4O15 (Ag/BaLa4Ti4O15)
exhibited the highest CO2 reduction performance, resulting in the formation of CO
as a main product. The Ag cocatalyst acted as a reduction site to produce CO from
CO2. Interestingly, the photocatalytic activity for CO2 reduction strongly depends
on the loading method of Ag nanoparticles. The small-size Ag nanoparticles
(10 nm) were prepared on the edge of plate particle of the BaLa4Ti4O15 by
liquid-phase chemical reduction method, and the performance of photocatalytic
CO2 reduction was significantly improved compare to the Ag/BaLa4Ti4O15 using
conventional method such as impregnation or photodeposition. It is concluded that
the separation of reaction sites (reduction and oxidation) and the loading of small-
size Ag cocatalyst on the edge of plate play important roles to achieve efficient
activation of CO2 reduction using Ag/BaLa4Ti4O15 photocatalyst.
As discussed in Sect. 9.11, the development of visible light-sensitive photocatalyst
is also absolutely imperative for more efficient CO2 reduction. The unique
photocatalyst system for CO2 reduction using composite oxide of Pt/ZnAl2O4 mod-
ified with mesoporous ZnGaNO was reported by Yan et al. [49]. The photocatalytic
reduction of CO2 to CH4 was efficiently confirmed in the presence of H2O under
visible light irradiation (λ > 420 nm), because this novel and unique photocatalyst
possesses high CO2 adsorption property derived from mesoporous structure, basicity,
and visible light absorption ability by narrow band gap structure to achieve efficient
CO2 reduction. This photocatalytic activity is about seven times higher than that of
Pt-loaded N-doped TiO2 known as versatile visible light-sensitive photocatalyst.

9.13 Localized Surface Plasmon Resonance-Enhanced


Photocatalyst

Au nanoparticle can absorb visible and infrared light in particular regions due to
localized surface plasmon resonance (LSPR) [50, 51]. In simple terms, LSPR is
made up of collective oscillations of free electrons in metal NPs driven by the
electromagnetic field of incident light [4]. This unique characteristic has given rise
to a new approach to the fabrication of visible light-sensitive photocatalysts, known
as plasmonic photocatalysts, through the deposition of Au nanoparticle onto suit-
able semiconductors [52]. It has been proposed that in the case of Au/TiO2, at an
excitation wavelength corresponding to the LSPR of Au, the Au nanoparticle
absorb photons and inject photogenerated electrons into the TiO2 conduction
band [53].
9 Reductive Conversion of Carbon Dioxide Using Various Photocatalyst Materials 241

Fig. 9.10 The energy level of anatase-type TiO2 and Au nanoparticle and reduction of CO2 to
HCHO, CH3OH, or CH4

The photocatalytic reduction of CO2 to CH4 using the unique LSPR property of
Au/TiO2 was recently performed by Hou et al. [54, 55]. In this system, the main
product was CH4, but the methanol and formaldehyde were also observed. When
the light (λ ¼ 532 nm) matched to the LSPR absorption of Au nanoparticle was
utilized in this system, the reduction efficiency of CO2 was about 24 times higher
than that of bare TiO2. They discussed that the excited electrons produced in the Au
nanoparticles by the LSPR did not directly transfer to the conduction band of TiO2.
However, when the photon possessing high enough energy to excite the electrons in
d-band of Au nanoparticles was irradiated, the charge was directly transferred
between Au nanoparticle and TiO2 (Fig. 9.10). Since the LSPR can absorb light
of long-wavelength region, this unique photocatalytic system is one of promising
approach for efficient CO2 reduction by beneficial use of photo-energy.

9.14 Conclusions

In this chapter, we have focused on the progress in the very stimulating field of the
photocatalytic reduction of CO2 with H2O on various types of Ti-oxide catalysts.
Upon UV-light irradiation, the reactions on TiO2 powders in the presence of gaseous
CO2 and H2O at 275 K produced CH4 as the major product, while on the highly
dispersed titanium oxide anchored on porous glass or ordered porous silicas, the
formations of CH3OH as well as CH4 were observed as the major products. In situ
spectroscopic studies of the system indicated that the photocatalytic reduction of CO2
with H2O is linked to a much higher activity of the charge-transfer excited state, i.e.,
(Ti3+—O)* of the tetrahedral coordinated Ti-oxide species formed on the surface.
Owing to the dramatically growing nature of this research field, we expect that the
242 K. Fuku et al.

design strategy described here can offer a helpful overview to the readers in this
fascinating area. In recent report, the highly efficient reduction of CO2 was achieved
using various oxide semiconductors or the LSPR property of metallic nanoparticles.
These photocatalytic systems are expected for solution of various global environment
problems including CO2 problem.

References

1. Anpo M, Yamashita H (1996) Photochemistry of surface species anchored on solid surface.


In: Anpo M (ed) Surface photochemistry. Wiley, Chichester, pp 117–164
2. Schiavello M (ed) (1997) Heterogeneous photocatalysts. Wiley, Chichester
3. Anpo M (2000) Photofunctional zeolites. NOVA, New York
4. Anpo M, Yamashita H, Zhang SG (1996) Photoinduced surface chemistry. Curr Opin Solid
State Mater Sci 1:630–635
5. Anpo M (1989) Photocatalysis on small particle TiO2 catalysts. Reaction intermediates and
reaction mechanisms. Res Chem Intermed 11:67–106
6. Anpo M, Ichihashi Y, Takeuchi M, Yamashita H (1998) Design of unique titanium oxide
photocatalysts by an advanced metal ion-implantation method and photocatalytic reactions
under visible light irradiation. Res Chem Intermed 24:143–149
7. Yamashita H, Ichihashi Y, Takeuchi M, Kishiguchi S, Anpo M (1999) Characterization of
metal ion-implanted titanium oxide photocatalysts operating under visible light irradiation.
J Synchrotron Radiat 6:451–452
8. Yamashita H, Honda M, Harada M, Ichihashi Y, Anpo M (1998) Preparation of titanium oxide
photocatalysts anchored on porous silica glass by a metal ion-implantation method and their
photocatalytic reactivities for the degradation of 2-propanol diluted in water. J Phys Chem B
102:10707–10711
9. Inoue T, Fujishima A, Konishi S, Honda K (1979) Photoelectrocatalytic reduction of carbon
dioxide in aqueous suspensions of semiconductor powders. Nature 277:637–640
10. Halmann M (1983) Photochemical fixation of carbon dioxide. In: Grätzel M (ed) Energy
resources through photochemistry and catalysis. Academic, New York, pp 507–565
11. Anpo M, Chiba K (1992) Photocatalytic reduction of carbon dioxide on anchored titanium
oxide catalysts. J Mol Catal 74:207–302
12. Yamashita H, Shiga A, Kawasaki S, Ichihashi Y, Ehara S, Anpo M (1995) Photocatalytic
synthesis of CH4 and CH3OH from CO2 and H2O on highly dispersed active titanium oxide
catalysts. Energy Conv Manag 36:617–620
13. Anpo M, Yamashita H, Ichihashi Y, Ehara S (1995) Photocatalytic reduction of CO2 with H2O
on various titanium oxide catalysts. J Electroanal Chem 396:21–26
14. Yamashita H, Nishiguchi H, Kamada N, Anpo M, Teraoka Y, Hatano H, Ehara S, Kikui K,
Palmisano L, Sclafani A, Schiavello M, Fox MA (1994) Photocatalytic reduction of CO2 with
H2O on TiO2 and Cu/TiO2 catalysts. Res Chem Intermed 20:815–823
15. Yamashita H, Kamada N, He H, Tanaka K, Ehara S, Anpo M (1994) Reduction of CO2 with
H2O on TiO2(100) and TiO2(110) single crystals under UV-irradiation. Chem Lett 1994
(5):855–858
16. Anpo M, Yamashita H, Ichihashi Y, Fujii Y, Honda M (1997) Photocatalytic reduction of CO2
with H2O on titanium oxides anchored within micropores of zeolites: effects of the structure of
the active sites and the addition of Pt. J Phys Chem B 101:2632–2636
17. Anpo M, Yamashita H, Fujii Y, Ichihashi Y, Zhang SG, Park DR, Ehara S, Park SE, Chang JS,
Yoo JW (1998) Photocatalytic reduction of CO2 with H2O on titanium oxides anchored within
zeolites. Stud Surf Sci Catal 114:177–182
9 Reductive Conversion of Carbon Dioxide Using Various Photocatalyst Materials 243

18. Yamashita H, Fuji Y, Ichihashi Y, Zhang SG, Ikeue K, Park DR, Koyano K, Tatsumi T,
Anpo M (1998) Selective formation of CH3OH in the photocatalytic reduction of CO2 with
H2O on titanium oxides highly dispersed within zeolites and mesoporous molecular sieves.
Catal Today 45:221–227
19. Anpo M, Zhang SG, Fujii Y, Ichihashi Y, Yamashita H, Koyano K, Tatsumi T (1998)
Photocatalytic reduction of CO2 with H2O on Ti-MCM-41 and Ti-MCM-48 mesoporous
zeolite catalysts. Catal Today 44:327–332
20. Ikeue K, Yamashita H, Anpo M (1999) Photocatalytic reduction of CO2 with H2O on titanium
oxides prepared within the FSM-16 mesoporous zeolite. Chem Lett 1999(11):1135–1136
21. Yamashita H, Ikeue K, Takewaki T, Anpo M (2002) In situ XAFS studies on the effects of the
hydrophobic-hydrophilic properties of Ti-beta zeolites in the photocatalytic reduction of CO2
with H2O. Top Catal 18:95–100
22. Yamashita H, Kawasaki S, Fujii Y, Ichihashi Y, Ehara S, Park SE, Chang JS, Yoo JW, Anpo M
(1998) Photocatalytic reduction of CO2 with H2O on Ti/Si binary oxide catalysts prepared by
the sol-gel method. Stud Surf Sci Catal 114:561–564
23. Saladin F, Forss L, Kamber I (1995) Photosynthesis of CH4 at a TiO2 surface from gaseous
H2O and CO2. J Chem Soc Chem Commun 5:533–534
24. Anpo M, Tomonari M, Fox MA (1989) In situ photoluminescence of titania as a probe of
photocatalytic reactions. J Phys Chem 93:7300–7303
25. Yamashita H, Ichihashi Y, Harada M, Stewart G, Fox MA, Anpo M (1996) Photocatalytic
degradation of 1-octanol on anchored titanium oxide and on TiO2 powder catalysts. J Catal
158:97–101
26. Frese KW (1991) Electrochemical reduction of carbon dioxide at intentionally oxidized copper
electrodes. J Electrochem Soc 138:3338–3343
27. Slamet HW, Nasution E, Purnama E, Kosela S, Gunlazuardi J (2005) Photocatalytic reduction
of CO2 on copper-doped titania catalysts prepared by improved-impregnation method. Catal
Commun 6:313–319
28. Raskö J, Solymosi F (1994) Infrared spectroscopic study of the photoinduced activation of
CO2 on TiO2 and Rh/TiO2 catalysts. J Phys Chem 98:7147–7152
29. Anpo M, Matsuoka M, Shioya Y, Yamashita H, Giamello E, Morterra C, Che M, Patterson
HH, Webber S, Ouellette S, Fox MA (1994) Preparation and characterization of the Cu+/ZSM-
5 catalyst and its reaction with NO under UV irradiation at 275 K. In situ photoluminescence,
EPR, and FT-IR investigations. J Phys Chem 98:5744–5750
30. Yamashita H, Matsuoka M, Tsuji K, Shioya Y, Anpo M, Che M (1996) In-situ XAFS,
photoluminescence, and IR investigations of copper ions included within various kinds of
zeolites. Structure of Cu(I) ions and their interaction with CO molecules. J Phys Chem
100:397–402
31. Yamashita H, Ichihashi Y, Anpo M, Hashimoto M, Louis C, Che M (1996) Cavities: the
structure and role of the active sites. J Phys Chem 100:16041–16044
32. Yamashita H, Zhang SG, Ichihashi Y, Matsumura Y, Souma S, Tatsumi T, Anpo M (1997)
Photocatalytic decomposition of NO at 275 K on titanium oxide catalysts anchored within
zeolite cavities and framework. Appl Surf Sci 121:305–309
33. Farges F, Brown GE Jr, Rehr JJ (1996) Coordination chemistry of Ti(IV) in silicate glasses and
melts: I. XAFS study of titanium coordination in oxide model compounds. Geochim
Cosmochim Acta 60:3023–3060
34. Anpo M, Aikawa N, Kubokawa Y, Che M, Louis C, Giamello E (1985) Photoluminescence
and photocatalytic activity of highly dispersed titanium oxide anchored onto porous Vycor
glass. J Phys Chem 89:5017–5021
35. Anpo M, Aikawa N, Kubokawa Y, Che M, Louis C, Giamello E (1985) Photoformation and
structure of oxygen anion radicals (O2) and nitrogen-containing anion radicals adsorbed on
highly dispersed titanium oxide anchored onto porous Vycor glass. J Phys Chem 89:5689–5694
36. Takewaki T, Hwang SJ, Yamashita H, Davis ME (1999) Synthesis of BEA-type molecular
sieves using mesoporous materials as reagents. Micropor Mesopor Mater 32:265–273
244 K. Fuku et al.

37. Camblor MA, Corma A, Esteve P, Martines M, Valencia S (1997) Epoxidation of unsaturated
fatty esters over large-pore Ti-containing molecular sieves as catalysts: important role of the
hydrophobic-hydrophilic properties of the molecular sieve. Chem Commun 33:795–796
38. Tatsumi T, Jappar N (1998) Properties of Ti-beta zeolites synthesized by dry-gel conversion
and hydrothermal methods. J Phys Chem B 102:7126–7131
39. Lin W, Han H, Frei H (2004) CO2 splitting by H2O to CO and O2 under UV light in TiMCM-41
silicate sieve. J Phys Chem B 108:18269–18273
40. Ikeue K, Nozaki S, Ogawa M, Anpo M (2002) Photocatalytic reduction of CO2 with H2O on
Ti-containing porous silica thin film photocatalysts. Catal Lett 80:111–114
41. Ikeue K, Nozaki S, Ogawa M, Anpo M (2002) Characterization of self-standing Ti-containing
porous silica thin films and their reactivity for the photocatalytic reduction of CO2 with
H2O. Catal Today 74:241–248
42. Yamashita H, Kawasaki S, Ichihashi Y, Harada M, Anpo M, Stewart G, Fox MA, Louis C,
Che M (1998) Characterization of titanium-silicon binary oxide catalysts prepared by the
sol-gel method and their photocatalytic reactivity for the liquid-phase oxidation of 1-octanol.
J Phys Chem B 102:5870–5875
43. Ozcan O, Yukruk F, Akkaya EU, Uner D (2007) Dye sensitized CO2 reduction over pure and
platinized TiO2. Top Catal 44:523–528
44. Nguyen TH, Wu JCS (2008) Photoreduction of CO2 to fuels under sunlight using optical-fiber
reactor. Sol Energy Mater Sol Cells 92:864–872
45. Navalon S, Dhakshinamoorthy A, Alvaro M, Garcia H (2013) You have free access to this
content photocatalytic CO2 reduction using non-titanium metal oxides and sulfides. Chem Sus
Chem 6:562–577
46. Xu Y, Schoonen MAA (2000) The absolute energy positions of conduction and valence bands
of selected semiconducting minerals. Am Mineral 85:543–556
47. Yahaya AH, Gondal MA, Hameed A (2004) Selective laser enhanced photocatalytic conver-
sion of CO2 into methanol. Chem Phys Lett 400:206–212
48. Iizuka K, Wato T, Miseki Y, Saito K, Kudo A (2011) Photocatalytic reduction of carbon
dioxide over Ag cocatalyst-loaded ALa4Ti4O15 (A ¼ Ca, Sr, and Ba) using water as a
reducing reagent. J Am Chem Soc 133:20863–20868
49. Yan S, Yu H, Wang N, Li Z, Zou Z (2012) Efficient conversion of CO2 and H2O into
hydrocarbon fuel over ZnAl2O4-modified mesoporous ZnGaNO under visible light irradiation.
Chem Commun 48:1048–1050
50. Mori K, Kawashima M, Che M, Yamashita H (2010) Enhancement of the photoinduced
oxidation activity of a ruthenium(II) complex anchored on silica-coated silver nanoparticles
by localized surface plasmon resonance. Angew Chem Int Ed 49:8598–8601
51. Fuku K, Hayashi R, Takakura S, Kamegawa T, Mori K, Yamashita H (2013) The synthesis of
size- and color-controlled silver nanoparticles by using microwave heating and their enhanced
catalytic activity by localized surface plasmon resonance. Angew Chem Int Ed 52:7446–7450
52. Cushing SK, Li JT, Meng FK, Senty TR, Suri S, Zhi MJ, Li M, Bristow AD, Wu NQ (2012)
Photocatalytic activity enhanced by plasmonic resonant energy transfer from metal to semi-
conductor. J Am Chem Soc 134:15033–15041
53. Tanaka A, Sakaguchi S, Hashimoto K, Kominami H (2012) Preparation of Au/TiO2 exhibiting
strong surface plasmon resonance effective for photoinduced hydrogen formation from organic
and inorganic compounds under irradiation of visible light. Catal Sci Technol 2:907–909
54. Hou WB, Hung WH, Pavaskar P, Goeppert A, Aykol M, Cronin SB (2011) Photocatalytic
conversion of CO2 to hydrocarbon fuels via plasmon-enhanced absorption and metallic
interband transitions. ACS Catal 1:929–936
55. Hou WB, Cronin SB (2013) A review of surface plasmon resonance-enhanced photocatalysis.
Adv Funct Mater 23:1612–1619
Chapter 10
Electrochemical Fixation of Carbon Dioxide

Hisanori Senboku

10.1 Introduction

Electrochemical fixation of carbon dioxide means an incorporation of carbon dioxide


in organic molecules as functional groups by the use of an electrochemical method.
Fixation of carbon dioxide can be principally classified into two categories. One is
fixation with C–C bond formation between carbon dioxide and organic molecules to
yield carboxylic acid. The electrochemical fixation of carbon dioxide in most reports
is in this category, and, consequently, electrochemical fixation of carbon dioxide
yielding carboxylic acid is often called “electrochemical carboxylation” or
“electrocarboxylation.” The other one is fixation of carbon dioxide with C–heteroatom
bond formation between a carbon atom of carbon dioxide and a heteroatom, such as an
oxygen or nitrogen atom, in organic molecules. As heteroatom sources, alcohol
and amine are used for this type of fixation to form carbonate and carbamate ions
(-X-CO2), respectively. Since the resulting carbonic acid and carbamic acid gener-
ally cannot be isolated due to their rapid decarboxylation, they are obtained and
isolated as their esters, carbonate and carbamate, respectively, by the reaction of the
forming carbonate or carbamate ions with appropriate electrophiles, such as alkyl
halides, in situ after the fixation of carbon dioxide (Scheme 10.1).
Of the two types of electrochemical fixation of carbon dioxide in organic
molecules, this chapter focuses on electrochemical fixation of carbon dioxide
with C–C bond formation giving carboxylic acid.
C–C bond formation between carbon dioxide and organic molecules is thought
to proceed generally via two different pathways. When the reduction potential of an
organic substrate is more positive than that of carbon dioxide, electrochemical
reduction of the organic substrate predominantly occurs to generate anionic species.
Nucleophilic attack of the resulting anionic species on carbon dioxide yields

H. Senboku (*)
Division of Chemical Process Engineering, Faculty of Engineering,
Hokkaido University, Sapporo, Hokkaido, Japan
e-mail: senboku@eng.hokudai.ac.jp

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 245
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_10,
© Springer-Verlag Berlin Heidelberg 2014
246 H. Senboku

i) fixation of carbon dioxide with C-C bond formation

organic + ne organic H3O organic


molecule FG molecule C C O molecule C C OH
CO2
O O
FG = functional groups carboxylic acid

ii) fixation of carbon dioxide with C-heteroatom bond formation

organic + ne organic R organic


molecule XH molecule X C O molecule X C OR
CO2
O O
X = O, NH X = O: carbonate
X = NH : carbamate

Scheme 10.1 Electrochemical fixation of carbon dioxide in organic molecules with C–C or
C–heteroatom bond formation

carboxylic acid (path a in Scheme 10.2). On the other hand, when the reduction
potential of the organic substrate is more negative than that of carbon dioxide,
one-electron reduction of carbon dioxide predominantly occurs to generate the
radical anion of carbon dioxide, which reacts with organic substrates, typically
alkenes, to yield the corresponding mono- and/or dicarboxylic acids (path b in
Scheme 10.2).
Fixation of carbon dioxide in organic molecules with C–C bond formation gives
carboxylic acid. Carbon dioxide used as a source of a carboxyl group is not only
abundant and economical but also nontoxic and attractive as an environmentally
benign C1 chemical reagent for organic synthesis, same as the description in other
chapters. By using an electrochemical method, efficient fixation of carbon dioxide
in appropriate organic molecules can be achieved even under an atmospheric
pressure of carbon dioxide under quite mild and almost neutral conditions. Elec-
trochemical reduction of organic molecules in the presence of carbon dioxide using
magnesium or aluminum as a sacrificial anode in a one-compartment electrochem-
ical cell has been found to be the most convenient, useful, and effective method
for efficient fixation of carbon dioxide to yield carboxylic acid in high yields
[1–4]. Carbon dioxide is also used for synthesizing carboxylic acid in conventional
methods. For example, reaction of a Grignard reagent or organolithium reagent
with carbon dioxide in ether or tetrahydrofuran, flammable organic solvents, gives
carboxylic acid. However, the reaction proceeds under highly basic conditions, and
the use of these highly reactive reagents imposes limitation of usable functional
groups in the reagents themselves and special caution in their handling is needed.
As well as carbon dioxide, a cyanide ion, such as KCN and NaCN, and carbon
monoxide (CO) are also effective C1 carbon sources in conventional methods for
synthesis of a nitrile and ester, which are the precursors of carboxylic acid and are
readily converted to carboxylic acid by hydrolysis. However, these C1 reagents
are unfortunately toxic, and from the viewpoint of green and sustainable chemistry,
10 Electrochemical Fixation of Carbon Dioxide 247

path a + 2e
R X R + X

R + CO2 R CO2

H3O
R CO2 R CO2 H

path b + e
CO2 CO2

+ CO2
CO2 R R

+e
CO2 CO2
R R

2 H3O CO2 H
R

CO2
R CO2 H
CO2 CO2 H
R
then
2 H3O

Scheme 10.2 Two general reaction pathways for electrochemical fixation of carbon dioxide with
C–C bond formation

the use of these reagents is unfavorable. On the other hand, an electrochemical


method is an environmentally benign method, and there is no need to use such toxic
reagents and flammable solvents with special caution in electrochemical fixation of
carbon dioxide to yield carboxylic acid in high yields. Although transition-metal
catalyzed fixation of carbon dioxide in organic and organometallic compounds has
recently been developed, an electrochemical method can achieve efficient fixation
of carbon dioxide without such expensive and air-sensitive metal catalysts by using
electrons as an essential reagent.
Platinum, grassy carbon, graphite, stainless steel, carbon fiber, silver, lead,
mercury pool, and some other metals have been reported to be usable as cathode
materials in electrochemical carboxylation. Among them, platinum, stainless steel,
carbon fiber, and graphite are frequently used for efficient formation of carboxylic
acid by electrochemical fixation of carbon dioxide with a sacrificial anode, such as
magnesium and aluminum [1–4], as a couple in an undivided cell (one-compartment
cell). DMF (N,N-dimethylformamide) and acetonitrile are mainly used for electro-
chemical carboxylation as solvents with low concentration of tetraalkylammonium
salts and lithium perchlorate as supporting electrolytes. Acetone, DMSO (dimethyl
sulfoxide), and THF (tetrahydrofuran) are also usable for fixation of carbon dioxide.
248 H. Senboku

10.2 Synthetic Applications

Efficient fixation of carbon dioxide in various kinds of organic molecules has been
successfully carried out by an electrochemical method with C–C bond formation;
alkenes [5] and activated alkenes including styrenes [5–15], allenes [5, 16], alkynes
[5, 17–19], enynes and diynes [5, 20–23], aromatic [1, 3, 24], allyl [3, 25, 26],
propargyl [27], benzyl [1, 3, 28–50] and vinyl [3, 51–53] halides, aryl [54, 55] and
vinyl triflates [56–60], aldehydes and ketones [61–64], imines [65], benzyl esters
and carbonates [66, 67], and other organic substrates [68–70] are efficiently
carboxylated by an electrochemical method to afford various kinds of useful
carboxylic acids [4, 71]. For example, electrochemical carboxylation of halides
1 and 3 gave fluorine-containing carboxylic acids 2 and ester 4 in high yields,
respectively [24, 34] (Scheme 10.3). Since synthesis of these carboxylic acids is
difficult by conventional method using organometallic reagents, it is obvious that
electrochemical method is of great use for synthesis of these carboxylic acids.
Electrochemical fixation of carbon dioxide into substituted allylic halide
5 was reported to occur at the α-position to give unsaturated carboxylic acid
6 regioselectively [25] while conventional reaction of the corresponding Grignard
reagent 7 derived from halide 5 with carbon dioxide occurs at the γ-position via
six-membered transition state to give regioisomer 8 as a sole product (Scheme 10.4).
Electrochemical debromination followed by carboxylation of 2-bromomethyl-
1,4-dibromobut-2-ene (9) can take place efficiently by using manganine, alloy
of 12–15 % Mn, 2–4 % Ni, and 80–85 % Cu as an anode metal to give 3-methy-
lene-4-pentenoic acid (10). This result shows that tribromide 9 can work as a
synthetic equivalent of isoprenyl carbanion 11. Carboxylic acid 10 has an isoprene
unit and can be used for aqueous Diels-Alder reaction as a water-soluble diene [26]
(Scheme 10.5).
Electrochemical carboxylation is sometimes carried out in the presence of a
transition-metal catalyst such as Pd [54–56, 72–74], Ni [5, 9, 16–23, 31, 32, 35, 38,
41, 52, 53, 60, 70, 75, 76], and cobalt [28, 36, 37, 44] complexes, the use of which

Cl CO2H
SUS Mg H
CF3 + CO2 CF3
DMF H2O
1 2 80–85%

CF3 CF2CO2Me
SUS Mg H
+ CO2
F DMF MeOH F
3 4 70%

Scheme 10.3 Synthesis of fluorine-containing carboxylic acids and esters by electrochemical


carboxylation of organic halides
10 Electrochemical Fixation of Carbon Dioxide 249

R2 Pt Mg H R2
+ CO2
R1 X DMF H2O R1 CO2H
5 6

R2 H R1 R 2
+ CO2
R1 MgX Et2O H2O CO2H
7 8

Scheme 10.4 Regioselective fixation of carbon dioxide in γ,γ-disubstituted allylic halides by


electrochemical method and Grignard reagents

Br
Pt MA
Br + CO2 CO2H
Br 0.1 M Bu4NI - DMF
9 2
10 mA/cm , 5 F/mol 10 57% CO2Me
-10 °C
NaHCO3-H2O
50 °C, 24h MeO2C
MA: manganine, alloy of 12–15% Mn, 2– 4% Ni, and 80– 85% Cu.

CO2Me
9
MeO2C
11
CO2H

Scheme 10.5 Synthesis of 3-methylene-4-pentenoic acid by electrochemical carboxylation of


2-bromomethyl-1,4-dibromobut-2-ene and its application as a water-soluble diene in aqueous
Diels-Alder reaction

often results in drastic enhancement and control of chemo- and regioselectivity and
efficiency of carbon dioxide fixation. For instance, electrochemical carboxylation
of lactone enol triflates 12, readily prepared from the corresponding lactones, in the
absence of any catalyst gives only a complex mixture in about 60 % conversion of
12, while similar carboxylation in the presence of 20 mol% of NiBr2•bpy results in
efficient fixation of carbon dioxide to give the corresponding cyclic α-alkoxyl-
α,β-unsaturated carboxylic acid 13 in high yields [60] (Scheme 10.6).
Electrochemical carboxylation of terminal alkynes 14 in the presence of Ni
catalyst gave 15 as a major product along with isomer 16 [17, 18]. Carboxylic
acids 17 and 18 are also synthesized by nickel-catalyzed electrochemical carbox-
ylation of terminal alkyne 14 [19]. Consequently, selective synthesis of each
carboxylic acid 15, 16, 17, and 18 from terminal alkyne 14 and carbon dioxide
can be accomplished even by selecting appropriate ligands on nickel(II) complexes
in the electrochemical carboxylation (Scheme 10.7). As a similar manner, control
250 H. Senboku

R1 O O R1 O OTf Pt Mg R1 O CO2H
+ CO2
( )n R2 ( )n R2 0.1 M Bu4NBF4 - DMF ( )n R2
10 mA/cm2, 3 F/mol
12 13 63 –79%
20 mol% NiBr2 bpy, 5 °C

Scheme 10.6 Synthesis of cyclic α-alkoxyl-α,β-unsaturated carboxylic acids by nickel(0)-


catalyzed electrochemical carboxylation of lactone enol triflates

H H
C Mg H R R
R C C H + CO2 H + CO2H
⬙Ni(II) cat - DMF H 2O CO2H H

14 15 16
R
CO2H R C C CO2H
CO2H
17 18

Scheme 10.7 Ligand-directed and nickel-catalyzed electrochemical carboxylation of terminal


alkynes

C Mg C Mg
+ CO2
CO2H Co(acac)2-PMTDA Ni(II) - TMEDA
DMF CO2H
20 19 21

Ni(II) - bpy
C Mg

CO2H

22

Scheme 10.8 Nickel-catalyzed electrochemical carboxylation of α,ω-diynes

of chemo- and regioselectivity by catalysts and ligands has also been achieved in
electrochemical carboxylation of α,ω-diyne 19 to give each carboxylic acids 20, 21,
and 22 selectively [22] (Scheme 10.8).
Carboxylation can take place with other electrochemically induced chemical
reactions such as cyclization [75–77]. For example, electrochemical carboxylation
of 2-allyloxybromobenzene (23) in the presence of nickel catalyst or electron transfer
10 Electrochemical Fixation of Carbon Dioxide 251

CO2, C Mg H
50% (24 : 25 = 5 : 1)
Ni(cyclam)(BF4)2 - DMF H2O
CO2H
X

O O
CO2, Pt Mg H
23: X = Br 24
25: X = CO2H DMF H 2O
82% (24 : 25 = 98 : 2)
t-Bu CO2Me

(electron transfer mediator)

Scheme 10.9 Synthesis of dihydrobenzofuran-3-ylacetic acids by electrochemical sequential


cyclization-carboxylation reaction

mediator results in cyclization followed by fixation of carbon dioxide to give


carboxylic acid 24 in high selectivity toward direct-carboxylation product 25
(Scheme 10.9) [75, 77].
One notable synthetic application of electrochemical fixation of carbon dioxide
is synthesis of 2-arylpropanoic acids, nonsteroidal anti-inflammatory drugs
(NSAIDs), and their useful precursors (Scheme 10.10). Electrochemical carboxyl-
ations of benzyl halide 26 give 2-arylpropanoic acid 27. Many successful syntheses
of NSAIDs, such as ibuprofen and naproxen, have been reported by this type of
reaction [1, 3, 30–33, 35–37, 40–44, 48–50, 71, 86, 87]. On the other hand, chiral
2-arylpropanoic acid 27 can be readily synthesized by electrochemical carboxyla-
tion of vinyl bromide 28 [51, 71] followed by enantioselective hydrogenation. For
example, electrochemical carboxylation of vinyl bromide 28a yields α, β-unsatu-
rated carboxylic acid 29a in 93 % yield [51], which can be readily transformed into
(S)-ibuprofen 27a by enantioselective hydrogenation [78, 79] (Scheme 10.10). The
precursors of naproxen (29b), ketoprofen (29c), and flurbiprofen (29c) are also
synthesized in high yields by similar electrochemical carboxylation. Consequently,
synthesis of chiral NSAIDs can be readily achieved from vinyl bromide 28 and
carbon dioxide only in two steps by electrochemical method (Scheme 10.11).
α,β-Unsaturated carboxylic acid 29 can also be obtained from aryl methyl ketone
30 by electrochemical carboxylation followed by dehydration of the resulting
α-hydroxycarboxylic acid 31. Successful syntheses of 31 as the precursors of chiral
NSAIDs have been reported by using this type of reaction [61, 63, 64, 71, 86, 87]
(Scheme 10.10).
It is noteworthy that syntheses of fluorinated analogs of NSAIDs can also be
achieved by electrochemical fixation of carbon dioxide [48–50]. For example,
electrochemical reduction of α,α-difluoroethylarenes 32 in the presence of
carbon dioxide by using a Pt cathode and a Mg anode resulted in the reductive
cleavage of one C–F bond followed by a fixation of carbon dioxide at the benzylic
252 H. Senboku

CH3 CH3
+ 2e, CO2
X CO2H
R R

26: X = Cl, Br 27
2-arylpropanoic acids

H2

+ 2e, CO2
Br CO2H
R R

28 29

- H2O

CH3 HO CH3
+ 2e, CO2
O CO2H
R R

30 31

Scheme 10.10 Three synthetic routes for nonsteroidal anti-inflammatory drugs (NSAIDs) having
2-arylpropanoic acid skeleton through electrochemical fixation of carbon dioxide

CH3
Br CO2, Pt Mg CO2H H2 CO2H
i-Bu 0.1 M Bu4NBF4 - DMF i-Bu Ru-(S)-BINAP i-Bu

28a 10 mA/cm2, 3 F/mol 29a 93% 27a: (S )-ibuprofen

O
CO2H
CO2H CO2H
MeO
F

29b: naproxen precursor (80%) 29c: ketoprofen precursor (74%) 29d: flurbiprofen precursor (75%)

Scheme 10.11 Efficient synthesis of 2-arylpropenoic acids, the precursors of chiral NSAIDs, by
electrochemical carboxylation of α-bromostyrene derivatives
10 Electrochemical Fixation of Carbon Dioxide 253

F F F CO2H
Pt Mg
CH3 + CO2 CH3
R R
0.1 M Bu4NBF4 - DMF
15 mA/cm2, 0 °C
32 33

F CO2H O F CO2H F CO2H


O
CH3 CH3 CH3
i-Bu

33a: α-fluoroibuprofen (82%) 33b: α-fluoroketoprofen (93%) 33c: α-fluoroloxoprofen (61%)

F F F CO2H
R Pt Mg R
CF3 + CO2 CF3
0.1 M Bu4NBF4 - DMF
15 mA/cm2, 6 F/mol, 0 °C
34 35

O F CO2H F CO2H
O
CF3 CF3

35a: tetrafluoroketoprofen (79%) 35b: tetrafluorofenoprofen (82%)

Scheme 10.12 Syntheses of mono- and tetrafluorinated NSAIDs by electrochemical carboxylation


of benzylic fluorides

position to give 2-fluoro-2-arylpropanoic acids, α-fluorinated NSAIDs, 33 in


good yields (Scheme 10.12). α-Fluoroibuprofen (33a), α-fluoroketoprofen (33b),
α-fluoroloxoprofen (33c), and other α-fluorinated NSAIDs are successfully synthe-
sized in high yields by similar electrochemical carboxylation [48]. Synthesis of
tetrafluorinated ketoprofen (35a) and fenoprofen (35b) is also accomplished using
pentafluoroethylarenes (34) as starting materials in similar electrochemical carbox-
ylation [49] (Scheme 10.12).
Electrochemical fixation of carbon dioxide is also an effective and powerful tool
for synthesis of β,β,β-trifluoro derivatives of ibuprofen, naproxen, and related
NSAIDs [50]. Electrochemical reduction of benzyl bromide 36, having
trifluoromethyl group at the benzylic position, in the presence of carbon dioxide
by using zinc, instead of magnesium, as an anode results in efficient fixation of
carbon dioxide to give the corresponding carboxylated product 37 in good yields.
Trifluoroibuprofen methyl ester (37a), trifluoronaproxen methyl ester (37b), and
trifluorofenoprofen (37c) are successfully synthesized in moderate to good yields
by the present electrochemical method (Scheme 10.13). Synthesis of 37 is quite
difficult by conventional chemical fixation of carbon dioxide using organometallic
reagents. Since β,β,β-trifluoroethyl anion such as 39 is mostly known to release a
254 H. Senboku

Br CO2H(Me)
Pt Zn
CF3 + CO2 CF3
R R
0.1 M Bu4NBF4 - DMF
(followed by TMSCHN2)
36 37

CO2Me CO2Me CO2H


O
CF3 CF3 CF3

i-Bu MeO

37a: trifluoroibuprofen 37b: trifluoronaproxen 37c: trifluorofenoprofen (80%)


methyl ester (42%) methyl ester (48%)

M M
- MF F
CF3 CF2
R R R
F F

38 39 40

Scheme 10.13 Synthesis of 2-aryl-3,3,3-trifluoropropanoic acids, trifluorinated NSAIDs, by


electrochemical carboxylation of (1-bromo-2,2,2-trifluoroethyl)arenes

CO2Et
Mg/ TMSCl / CO2 EtOH
CO2Et CO2Et
Ar Ar
DMF dioxane H2SO4
41 42 42–71%yields

Scheme 10.14 Magnesium-promoted carboxylation of ethyl cinnamates

fluoride ion spontaneously to give difluoroalkenes such as 40 [50, 80], it is hard to


prepare the corresponding organometallic reagents such as 38 by conventional
chemical methods.
These successful syntheses of fluorinated NSAIDs 33, 35, and 37 strongly
indicate the usefulness, significance, advantage, and convenience of an electro-
chemical fixation of carbon dioxide for synthesis of carboxylic acids.
Efficient fixation of carbon dioxide can also be achieved by using magnesium
metal, instead of electrolysis, as an electron donor for the reduction of organic
substrate. For example, magnesium-promoted carboxylation of ethyl cinnamate
derivatives 41 for practical synthesis of arylsuccinate 42 has been successfully
carried out [81] (Scheme 10.14).
It is well known that carbon dioxide can become supercritical fluid under
relatively moderate conditions (Tc ¼ 31 C, Pc ¼ 7.5 MPa). Supercritical carbon
dioxide (scCO2) has significant potential as an environmentally benign solvent to
replace the hazardous organic solvents, as carbon dioxide is abundant, economical,
10 Electrochemical Fixation of Carbon Dioxide 255

Pt Mg
Ar X + scCO2 Ar CO2H
Bu4NBF4 - CH3CN
43 25 mA/cm2, 3 F/mol 44
[40 °C, 80 kg/cm2]

X
X
X
AcO

43a: X = Br 43b: X = Br 43c: X = Br


86% 92% 74%
44a: X = CO2H 44b: X = CO2H 44c: X = CO2H

Scheme 10.15 Electrochemical carboxylation of aryl halides in supercritical carbon dioxide


(scCO2)

and nontoxic and can be recovered and reused after the reaction. The use of scCO2
to electrochemical carboxylation has been studied, and a novel method for electro-
chemical carboxylation in scCO2 by the use of small amount of acetonitrile (5–10 to
155 mL of scCO2) as a cosolvent has been developed [82]. No current flows in pure
supercritical carbon dioxide due to its poor conductivity.
Fixation of carbon dioxide to aromatic compounds successfully takes place in
scCO2. Electrochemical reduction of aromatic halides 43 in scCO2 containing a
small amount of acetonitrile with a platinum plate cathode and a magnesium anode
results in efficient fixation of carbon dioxide to give aryl carboxylic acids such as
44a–c in high yields [82, 83] (Scheme 10.15).
Electrochemical dicarboxylation of phenanthrene (45) and anthracene (47) also
proceeds efficiently in scCO2 to yield the corresponding dicarboxylic acids 46 and
48, respectively, in high stereoselectivity and high yields [83, 84] (Scheme 10.16).
Similar electrochemical dicarboxylation of 47 in acetonitrile solution in the pres-
ence of atmospheric pressure of carbon dioxide only gives 32 % yield of dicarbox-
ylic acid 48.
Electrochemical carboxylation in scCO2 can also be applied to the synthesis of
NSAIDs. For example, electrolysis of benzyl chloride 49 in scCO2 yields racemic
naproxen (50) in 74 % yield [83, 84] (Scheme 10.17).
Electrochemical carboxylation of aryl methyl ketone 51 in scCO2 gives
α-hydroxycarboxylic acid 52 in 78 % yield (Scheme 10.18), which can be transformed
into chiral (S)-naproxen (53) by dehydration followed by enantioselective hydroge-
nation (Scheme 10.18; see also Scheme 10.10). Synthesis of loxoprofen precursor 55
is of considerable synthetic as well as electrochemical significance as the usual
chemical reaction cannot control the reactivity of the two carbonyl groups in 54. In
the electrochemical reactions, however, aryl methyl ketone is more readily reduced
than cyclopentanone, and the electrochemical carboxylation of 54 selectively takes
place at the aryl ketone moiety to give expected α-hydroxypropanoic acid 55 in 91 %
256 H. Senboku

HO2C CO2H

Pt Mg
+ scCO2
Bu4NBF4 - CH3CN
25 mA/cm2, 3 F/mol
45 46 93%
[40 °C, 80 kg/cm2]

CO2H

Pt Mg
+ scCO2
Bu4NBF4 - CH3CN
25 mA/cm2, 3 F/mol
CO2H
[40 °C, 80 kg/cm2]
47 48 93%

Scheme 10.16 Electrochemical carboxylation of phenanthrene and anthracene in scCO2

Cl Pt Mg CO2H
+ scCO2
MeO Bu4NBF4 - CH3CN MeO
20 mA/cm2, 5 F/mol
49 [40 °C, 80 kg/cm2] 50: racemic naproxen 74%

Scheme 10.17 Synthesis of naproxen by electrochemical carboxylation in scCO2

yield [84], which can also be transformed into chiral loxoprofen (56) by dehydration
followed by enantioselective hydrogenation (Scheme 10.18, see also Scheme 10.10).
Similar electrochemical carboxylation of vinyl bromides 28 in scCO2 also gives
the precursor of NSAIDs 29a, 29b, and 29d in almost the same yields as those using
atmospheric pressure of carbon dioxide [84] (see Scheme 10.11).
Although there are many reports on electrochemical carboxylation, only a
few studies on diastereoselective [85, 86] and asymmetric [87, 88] fixation of
carbon dioxide by electrochemical method have been reported. For example,
diastereoselective electrochemical carboxylation using chiral N-(2-bromoacyl)
oxazolidin-2-ones as substrates has been studied, and the best diastereomeric
ratio (58/59 ¼ 2/98) with 80 % isolated yield is accomplished when amide 57
(R ¼ Me) having Oppolzer’s camphorsultam as chiral auxiliary is used as a
substrate in electrochemical carboxylation [85] (Scheme 10.19).
Alkaloid-induced asymmetric electrochemical carboxylation of aryl alkyl ketone
has been studied. Yields and enantiomeric excess of carboxylic acids are, however,
still moderate, and further development of these areas including diastereoselective
fixation of carbon dioxide with wide variety of examples and scopes are desirable.
10 Electrochemical Fixation of Carbon Dioxide 257

O HO CO2H
scCO2, Pt Mg CO2H
Bu4NBF4 - CH3CN MeO MeO
MeO
20 mA/cm2, 5 F/mol
51 52 78% 53: (S)-naproxen
[40 °C, 80 kg/cm2]

O HO CO2H
O O O
scCO2, Pt Mg CO2H
Bu4NBF4 - CH3CN
54 20 mA/cm2, 5 F/mol 55 91% 56: chiral loxoprofen
[40 °C, 80 kg/cm2]

Scheme 10.18 Synthesis of the precursors of naproxen and loxoprofen by electrochemical


carboxylation of aryl methyl ketones in scCO2

Pt Al
O + CO2 O + O
Bu4NBF4 - THF
R N S 4 mA/cm2, 3 F/mol R N S R N S
O O O
Br O -20 °C MeO2C O MeO2C O
57 then CH2N2 58 59
R = Me, Et, Bu, i-Pr 70–80% yields

58 / 59 = 2–4 / 98–96

Scheme 10.19 Diastereoselective electrochemical carboxylation of chiral α-bromoamides

The use of an ionic liquid as a solvent and a supporting electrolyte for


electrochemical fixation of carbon dioxide has been studied [89, 90]. Since ionic
liquid can be readily recovered and reused after the reaction, it has significant
possibility as a novel reaction medium not only for fixation of carbon dioxide
but also for other electrochemical reactions. As well as diastereoselective and
asymmetric electrochemical fixation of carbon dioxide, development of these
research areas are now in progress and further studies are also expected.

10.3 Conclusion

The present electrochemical method for efficient fixation of atmospheric pressure of


carbon dioxide to a variety of organic compounds yielding useful carboxylic acids
has several advantages: use of simple one-compartment cell, simple electrolysis
under constant current conditions, high yields of the produced carboxylic acid
when a sacrificial anode such as magnesium and aluminum is used, and easy
application to the synthesis of useful precursors of nonsteroidal anti-inflammatory
258 H. Senboku

agents such as ibuprofen and naproxen and other high value-added substances.
This electrochemical method might be used for industrial production of these com-
pounds. The use of supercritical carbon dioxide for the electrochemical fixation of
carbon dioxide would become a useful method for production of carboxylic acids in
the future because the use of organic solvent will be highly limited and both
environmentally benign procedures and recycle of carbon source should be important
and necessary in organic synthesis from the viewpoint of green and sustainable
chemistry.

References

1. Silvestri G, Gambino S, Filardo G, Gulotta A (1984) Sacrificial anodes in the electrocar-


boxylation of organic chlorides. Angew Chem Int Ed Engl 23:979–980
2. Silvestri G, Gambino S, Filardo G (1991) Use of sacrificial anodes in synthetic electrochem-
istry. Processes involving carbon dioxide. Acta Chem Scand 45:987–992
3. Sock O, Troupel M, Périchon J (1985) Electrosynthesis of carboxylic acids from organic
halides and carbon dioxide. Tetrahedron Lett 26:1509–1512
4. Chaussard J, Folest JC, Nédélec JY, Périchon J, Sibille S, Troupel M (1990) Use of sacrificial
anodes in electrochemical functionalization of organic halides. Synthesis 1990:369–381
5. Dérien S, Clinet J-C, Duñach E, Péricon J (1992) Electrochemical incorporation of carbon
dioxide into alkenes by nickel complexes. Tetrahedron 48:5235–5248
6. Tyssee DA, Baiser MM (1974) Electrocarboxylation I. Mono- and dicarboxylation of activated
olefins. J Org Chem 39:2819–2823
7. Harada J, Sakakibara Y, Kunai A, Sasaki K (1984) Electrochemical carboxylation of α,
β-unsaturated ketones with carbon dioxide. Bull Chem Soc Jpn 57:611–612
8. Filardo G, Gambino S, Silvestri G, Gennaro A, Vianello E (1984) Electrocarboxylation of
styrene through homogeneous redox catalysis. J Electroanal Chem 177:303–309
9. Chiozza E, Desigaud M, Greiner J, Duñach E (1998) First example of double bond migration
in the electrochemical CO2 incorporation into (perfluoroalkyl)alkenes. Tetrahedron Lett
39:4831–4834
10. Ballivet-Tkatchenko D, Folest JC, Tanji J (2000) Electrocatalytic reduction of CO2 for the
selective carboxylation of olefins. Appl Organometal Chem 14:847–849
11. Senboku H, Komatsu H, Fujimura Y, Tokuda M (2001) Efficient electrochemical
dicarboxylation of phenyl-substituted alkenes: synthesis of 1-phenylalkane-1,2-dicarboxylic
acids. Synlett 2001(3):418–420
12. Chowdhury MA, Senboku H, Tokuda M (2004) Electrochemical carboxylation of bicycle
[n.1.0]alkylidene derivatives. Tetrahedron 60:475–481
13. Senboku H, Yamauchi Y, Kobayashi N, Fukui A, Hara S (2011) Electrochemical carboxyla-
tion of flavones: facile synthesis of flavanone-2-carboxylic acids. Electrochemistry
79:862–864
14. Senboku H, Yamauchi Y, Kobayashi N, Fukui A, Hara S (2012) Some mechanistic studies on
electrochemical carboxylation of flavones to yield flavanone-2-carboxylic acids. Electrochim
Acta 82:450–456
15. Gambino S, Filardo G, Silvestri G (1982) Electrochemical carboxylation of organic substrates.
Synthesis of carboxylic derivatives of acenaphthylene. J Appl Electrochem 12:549–555
16. Dérien S, Clinet J-C, Duñach E, Péricon J (1990) Coupling of allenes and carbon dioxide
catalyzed by electrogenerated nickel complexes. Synlett:361–364
17. Duñach E, Péricon J (1988) Electrochemical carboxylation of terminal alkynes catalyzed by
nickel complexes: unusual regioselectivity. J Organometal Chem 352:239–246
10 Electrochemical Fixation of Carbon Dioxide 259

18. Duñach E, Péricon J (1990) New catalytic system for the activation of carbon dioxide and
alkynes based on electrogenerated nickel complexes. Synlett 1990(3):143–145
19. Labbé E, Duñach E, Péricon J (1988) Ligand-directed reaction products in the nickel-catalyzed
electrochemical carboxylation of terminal alkynes. J Organometal Chem 353:C51–C56
20. Dérien S, Clinet JC, Duñach E, Péricon J (1992) New C-C bond formation through the nickel-
catalysed electrochemical coupling of 1,3-enynes and carbon dioxide. J Organometal
Chem 424:213–224
21. Dérien S, Clinet JC, Duñach E, Péricon J (1993) Activation of carbon dioxide: nickel-
catalyzed electrochemical carboxylation of diynes. J Org Chem 58:2578–2588
22. Dérien S, Duñach E, Péricon J (1990) Electrogenerated nickel(0) catalyzed carbon dioxide
incorporation into α, ω-diynes. J Organometal Chem 385:C43–C46
23. Dérien S, Clinet JC, Duñach E, Péricon J (1991) First example of direct carbon dioxide
incorporation into 1,3-diynes: a highly regio- and stereo-selective nickel-catalyzed electro-
chemical reaction. J Chem Soc Chem Commun:549–550
24. Heintz M, Sock O, Saboureau C, Périchon J, Troupel M (1988) Electrosynthesis of aryl-
carboxylic acids from chlorobenzene derivatives and carbon dioxide. Tetrahedron 44:1631–1636
25. Tokuda M, Kabuki T, Kato Y, Suginome H (1995) Regioselective synthesis of β, γ-unsaturated
carboxylic acids by the electrochemical carboxylation of allylic bromides using a reactive-
metal anode. Tetrahedron Lett 36:3345–3348
26. Tokuda M, Yoshikawa A, Suginome H, Senboku H (1997) New and convenient synthesis of
3-methylenepent-4-enoic acid by electrochemical carboxylation. Synthesis 1997(10):1143–1145
27. Tokuda M, Kabuki T, Suginome H (1994) Electrochemical carboxylation of propargylic halides
by the use of reactive-metal anode. Denki Kagaku (presently Electrochemistry) 62:1144–1147
28. Folest JC, Duprilot JM, Périchon J, Robin Y, Devynck J (1985) Electrocatalyzed carboxylation
of organic halides by a cobalt-salen complex. Tetrahedron Lett 26:2633–2636
29. Silvestri G, Gambino S, Filardo G, Greco G, Gulotta A (1986) Electrochemical carboxylation
of benzal chloride. Tetrahedron Lett 25:4307–4308
30. Fauvarque JF, Jutand A, Francois M (1986) Electrosynthesis of 2-arylpropionic acids from
ArCHMeCl and carbon dioxide, catalyzed by nickel complexes. Synthesis of anti-
inflammatory agents. Nouv J Chim 10:119–122
31. Fauvarque JF, Jutand A, Francois M (1988) Nickel catalysed electrosynthesis of anti-
inflammatory agents. Part I – Synthesis of aryl-2 propionic acids, under galvanostatic condi-
tions. J Appl Electrochem 18:109–115
32. Fauvarque JF, Jutand A, Francois M, Petit MA (1988) Nickel-catalyzed electrosynthesis of
anti-inflammatory agents: Part II: Monitoring of the electrolyses by HPLC analysis: role of the
catalyst. J Appl Electrochem 18:116–119
33. Chaussard J, Troupel M, Robin Y, Jacob G, Juhasz JP (1989) Scale-up of electrocarboxylation
reactions with a consumable anode. J Appl Electrochem 19:345–348
34. Saboureau C, Troupel M, Sibille S, Périchon J (1989) Electroreductive coupling of
trifluoromethyl arenes with electrophiles: synthetic applications. J Chem Soc Chem
Commun:1138–1139
35. Fauvarque JF, De Zelicourt Y, Amatore C, Jutand A (1990) Nickel-catalyzed electrosynthesis
of anti-inflammatory agents: Part III: A new electrolyzer for organic solvents: oxidation
of metal power as an alternative sacrificial anodes. J Appl Electrochem 20:338–340
36. Isse AA, Gennaro A, Vianello E (1996) Electrochemical carboxylation of aryl methyl chlo-
rides catalysed by [Co(salen)][H2 salen ¼ N,N0 -bis(salicylidene)ethane-1,2-diamine]. J Chem
Soc Dalton Trans:1613–1618
37. Chung WH, Guo P, Wong CP, Lau CP (2000) Electrocarboxylation of aryl methyl chlorides
catalyzed by cobalt 6,60 -bis(2-hydroxyphenyl)-2,20 -bipyridine. J Electroanal Chem 486:32–39
38. Gennaro A, Isse AA, Maran F (2001) Nickel(I)(salen)-electrocatalyzed reduction of benzyl
chloride in the presence of carbon dioxide. J Electroanal Chem 507:124–134
39. Isse AA, Gennaro A (2002) Electrocatalytic carboxylation of benzyl chlorides at silver
cathodes in acetonitrile. Chem Commun:2798–2799
260 H. Senboku

40. Stepanov AA, Volodin YY, Grachev MK, Kurochkina GI, Syrtsev AN, Grinberg VA (2002)
Reactions of direct electrocarboxylation of 1-phenylethylbromide and 1-(4-isobutylphenyl)
ethyl chloride in the presence of β-cyclodextrin. Russ J Electrochem 38:1346–1352
41. Isse AA, Ferlin MG, Gennaro A (2003) Homogeneous electron transfer catalysis in the
electrochemical carboxylation of arylethyl chlorides. J Electroanal Chem 541:93–101
42. Isse AA, Ferlin MG, Gennaro A (2005) Electrocatalytic reduction of arylethyl chlorides at
silver cathodes in the presence of carbon dioxide: synthesis of 2-arylpropanoic acids.
J Electroanal Chem 581:38–45
43. Damodar J, Krishna Mohan S, Khaja Lateef S, Jayarama Reddy S (2005) Electrosynthesis of
2-arylpropionic acids from α-methylbenzyl chlorides and carbon dioxide by [Co(Salen)].
Synth Commun 35:1143–1150
44. Ramesh Raju R, Khaja Lateef S, Krishna Mohan S, Jayarama Reddy S (2006) Synthesis of aryl-
2-propionic acids by electrocarboxylation of methyl aryl bromides. Arkivoc 2006(i):76–80
45. Isse AA, De Giusti A, Gennaro A, Falciola L, Mussini PR (2006) Electrochemical reduction of
benzyl halides at a silver electrode. Electrochim Acta 51:4956–4964
46. Scialdone O, Galia A, Errante G, Isse AA, Gennaro A, Filardo G (2008) Electrocarboxylation of
benzyl chlorides at silver cathode at the preparative scale level. Electrochim Acta 53:2514–2528
47. Niu DF, Xiao LP, Zhang AJ, Zhang GR, Tan QY, Lu JX (2008) Electrocatalytic carboxylation
of aliphatic halides at silver cathode in acetonitrile. Tetrahedron 64:10517–10520
48. Yamauchi Y, Fukuhara T, Hara S, Senboku H (2008) Electrochemical carboxylation of α,
α-difluorotoluene derivatives and its application to the synthesis of α-fluorinated nonsteroidal
anti-inflammatory drugs. Synlett 2008:438–442
49. Yamauchi Y, Sakai K, Fukuhara T, Hara S, Senboku H (2009) Synthesis of 2-aryl-2,3,3,3-
tetrafluoropropanoic Acids, tetrafluorinated fenoprofen and ketoprofen by electrochemical
carboxylation of pentafluoroethylarenes. Synthesis:3375–3377
50. Yamauchi Y, Hara S, Senboku H (2010) Synthesis of 2-aryl-3,3,3-trifluoropropanoic acids
using electrochemical carboxylation of (1-bromo-2,2,2-trifluoroethyl)arenes and its application
to the synthesis of β, β, β-trifluorinated non-steroidal anti-inflammatory drugs. Tetrahedron
2009(66):473–479
51. Kamekawa H, Senboku H, Tokuda M (1997) Facile synthesis of aryl-substituted 2-alkenoic
acids by electroreductive carboxylation of vinylic bromides using a magnesium anode.
Electrochim Acta 42:2117–2123
52. Kamekawa H, Kudoh H, Senboku H, Tokuda M (1997) Synthesis of α, β-unsaturated carbox-
ylic acids by nickel(II)-catalyzed electrochemical carboxylation of vinyl bromides. Chem Lett
26:917–918
53. Kuang C, Yang Q, Senboku H, Tokuda M (2005) Stereospecific electrochemical carboxylation
of β-bromostyrene by use of nickel(II) catalyst. Chem Lett 34:528–529
54. Jutand A, Négri S, Mosleh A (1992) Palladium catalyzed electrosynthesis using aryl trifluor-
omethanesulfonates (triflates). Synthesis of biaryls and aromatic carboxylic acids. J Chem Soc
Chem Commun:1729–1730
55. Jutand A, Négri S (1998) Activation of aryl and vinyl triflates by palladium and electron
transfer–electrosynthesis of aromatic and α, β-unsaturated carboxylic acids from carbon
dioxide. Eur J Org Chem 1998:1811–1812
56. Jutand A, Négri S (1997) Palladium-catalyzed carboxylation of vinyl triflates. Electrosynthesis
of α, β-unsaturated carboxylic acids. Synlett 1997:719–721
57. Kamekawa H, Senboku H, Tokuda M (1998) New electrochemical carboxylation of vinyl
triflates. Synthesis of β-keto carboxylic acid. Tetrahedron Lett 39:1591–1594
58. Senboku H, Fujimura Y, Kamekawa H, Tokuda M (2000) Divergent electrochemical carbox-
ylation of vinyl triflates: new electrochemical synthesis of phenyl-substituted α, β-unsaturated
carboxylic acids and aliphatic β-keto carboxylic acids. Electrochim Acta 45:2995–3003
59. Senboku H, Kanaya H, Fujimura Y, Tokuda M (2001) Stereochemical study on electrochemical
carboxylation of vinyl triflates. J Electroanal Chem 507:82–88
10 Electrochemical Fixation of Carbon Dioxide 261

60. Senboku H, Kanaya H, Tokuda M (2002) Convenient synthesis of cyclic α-alkoxyl-α,


β-unsaturated carboxylic acids by nickel-catalyzed electrochemical carboxylation of lactone
enol triflates. Synlett 2002:140–142
61. Silvestri G, Gambino S, Filardo G (1986) Electrochemical carboxylation of aldehydes and
ketones with sacrificial aluminum anodes. Tetrahedron Lett 27:3429–3430
62. Mcharek S, Heintz M, Troupel M, Péricon J (1989) Electrocarboxylation de composés
carbonyles aliphatiques, aromatiques et vinyliques: intérêt de l’utilisation d’une anode con-
sumable en magnésium. Bull Chim Soc Fra:95–97
63. Chan ASC, Huang TT, Wagenknecht JH, Miller RE (1995) A novel synthesis of 2-arylacetic
acids via electrocarboxylation of methyl aryl ketones. J Org Chem 60:742–744
64. Datt AK, Marron PA, King CJH, Wagenknecht JH (1998) Process development for electrocar-
boxylation of 2-acetyl-6-methoxynaphthalene. J Appl Electrochem 28:569–577
65. Silvestri G, Gambino S, Filardo G (1988) Synthesis of N-substituted α-amino acids by
electrocarboxylation of imines with sacrificial metallic anodes. Gazz Chim Ital 118:643–648
66. Gal J, Folest JC, Troupel M, Moingeon MO, Chaussard J (1995) Synthesis of arylacetic and
2-arypropionic acids by electrocarboxylation of benzylic compounds bearing a leaving group
other than a halogen. New J Chem 19:401–407
67. Ohkoshi M, Michinishi J, Hara S, Senboku H (2010) Electrochemical carboxylation of
benzylic carbonates: alternative method for efficient synthesis of arylacetic acids. Tetrahedron
66:7732–7737
68. Pokhodenko VD, Koshechko VG, Titov VE, Lopushanskaja VA (1995) A convenient elec-
trochemical synthesis of α-oxoacids. Tetrahedron Lett 36:3277–3278
69. Koshechko VG, Titov VE, Lopushanskaya VA (2002) Electrochemical carboxylation of
benzoyl bromides as effective phenylglyoxylic acid synthesis route. Electrochem Commun
4:655–658
70. Medeiros MJ, Pintaric C, Olivero S, Duñach E (2011) Nickel-catalyzed electrochemical
carboxylation of allylic acetates and carbonates. Electrochim Acta 56:4384–4389
71. Tokuda M (2006) Efficient fixation of carbon dioxide by electrolysis – facile synthesis of
useful carboxylic acids. J Nat Gas Chem 15:275–281
72. Torii S, Tanaka H, Hamatani T, Morisaki K, Jutand A, Pfluger F, Fauvarque JF (1986) Pd(0)-
catalyzed electroreductive carboxylation of aryl halides, β-bromostyrene, and allyl acetates
with CO2. Chem Lett 15:169–172
73. Amatore C, Jutand A, Khalil F, Nielsen MF (1992) Carbon dioxide as a C1 building block.
Mechanism of palladium-catalyzed carboxylation of aromatic halides. J Am Chem Soc
114:7076–7085
74. Damodar J, Krishna Mohan S, Jayarama Reddy S (2001) Synthesis of 2-arylpropionic acids by
electrocarboxylation of benzyl chlorides catalysed by PdCl2(PPh3)2. Electrochem Commun
3:762–766
75. Olivero S, Dunach E (1999) Selectivity in the tandem cyclization – carboxylation reaction of
unsaturated haloaryl ethers catalyzed by electrogenerated nickel complexes. Eur J Org Chem
1999:1885–1891
76. Tyssee DA, Baiser MM (1974) Electrocarboxylation II. Electrocarboxylative dimerization and
cyclization. J Org Chem 39:2823–2828
77. Senboku H, Michinishi J, Hara S (2011) Facile synthesis of 2,3-dihydrobenzofuran-3-ylacetic
acids by novel electrochemical sequential aryl radical cyclization-carboxylation of
2-allyloxybromobenzenes using methyl 4-tert-butylbenzoate as an electron-transfer mediator.
Synlett 2011:1567–1572
78. Manimaran T, Wu TC, Klobucar WD, Kolich CH, Stahly GP, Fronczek FR, Watkins SE
(1993) In situ generation of ruthenium-chiral phosphine complexes and their use in asymmet-
ric hydrogenation. Organometallics 12:1467–1470
79. Zhang X, Uemura T, Matsumura K, Sayo N, Kumobayashi H, Takaya H (1994) Highly
enantioselective hydrogenation of α, β-unsaturated carboxylic acids catalyzed by
H8-BINAP-Ru(II) complexes. Synlett 1994:501–503
262 H. Senboku

80. Uneyama K, Katagiri T, Amii H (2008) α-Trifluoromethylated carbanion synthons. Acc Chem
Res 41:817–829
81. Maekawa H, Murakami T, Miyazaki T, Nishiguti I (2011) Practical synthesis of diethyl
phenylsuccinate by Mg-promoted carboxylation of ethyl cinnamate. Chem Lett 40:368–369
82. Sasaki A, Kudoh H, Senboku H, Tokuda M (1998) Electrochemical carboxylation of several
organic halides in supercritical carbon dioxide. In: Torii S (ed) Novel trends in electroorganic
synthesis. Springer, Tokyo, pp 245–246
83. Tokuda M (1999) Electrolytic synthesis using supercritical carbon dioxide. Electrochemistry
67:993–996
84. Senboku H, Tokuda M (2002) Electrochemical organic synthesis in supercritical carbon
dioxide. Fine Chemical 31:50–60
85. Feroci M, Orsini M, Palombi L, Sotgiu G, Colapietro M, Ineci A (2004) Diastereoselective
electrochemical carboxylation of chiral α-bromocarboxylic acid derivatives: an easy access to
unsymmetrical alkylmalonic ester derivatives. J Org Chem 69:487–494
86. Orsini M, Feroci M, Sotgiu G, Inesi A (2005) Stereoselective electrochemical carboxylation:
2-phenylsuccinates from chiral cinnamic acid derivatives. Org Biomol Chem 3:1203–1208
87. Zhang K, Wang H, Zhao S, Niu D, Lu J (2009) Asymmetric electrochemical carboxylation of
prochiral acetophenone: an efficient route to optically active atrolactic acid via selective
fixation of carbon dioxide. J Electroanal Chem 630:35–41
88. Zhao S, Zhu M, Zhang K, Wang H, Lu J (2011) Alkaloid induced asymmetric electrochemical
carboxylation of 4-methylpropiophenone. Tetrahedron Lett 52:2702–2705
89. Wang H, Zhang G, Liu Y, Luo Y, Lu J (2007) Electrocarboxylation of activated olefins in ionic
liquid BMIMBF4. Electrochem Commun 9:2235–2239
90. Hiejima Y, Hayashi M, Uda A, Oya S, Kondo H, Senboku H, Takahashi K (2010) Electro-
chemical carboxylation of α-chloroethylbenzene in ionic liquids compressed with carbon
dioxide. Phys Chem Chem Phys 12:1953–1957
Chapter 11
Accelerated Carbonation Processes
for Carbon Dioxide Capture, Storage
and Utilisation

Renato Baciocchi, Giulia Costa, and Daniela Zingaretti

11.1 Introduction

This chapter deals with the application of accelerated carbonation of alkaline minerals
and industrial residues in the framework of the actions aimed at reducing carbon
dioxide emissions to the atmosphere. Depending on the point of view and the type of
application, carbonation can be seen as either a CO2 capture, CO2 storage or CO2
utilisation process. The boundary between the three definitions is however not clear-
cut and depends on the main environmental benefit brought by the implementation of
the carbonation process. The definition of carbonation as a utilisation or storage option
will eventually depend on the extent of storage capacity that we may be able to achieve
with respect to the emissions of the process we are dealing with. For instance, if we
consider the case of accelerated carbonation of alkaline industrial residues, we may
consider this process either as a CO2 storage or a CO2 utilisation one; this will depend
on the relevance we may give to the CO2 uptake capacity of these materials or to the
improvement of their environmental properties after carbonation. Although this
chapter is part of a book on CO2 utilisation, it is worth pointing out that, differently
from the other utilisation options considered, carbonation allows for a net storage of
CO2 into a solid and thermodynamically stable form and thus may be considered as a
truly effective measure for climate change mitigation.
This chapter will first address the fundamentals of carbonation, describing the
thermodynamics and kinetics of the process with reference to suitable raw materials
and minerals. This will lead to discuss naturally occurring carbonation, highlighting
the main issues that hinder carbonation to take place extensively on the earth’s crust
and possibly constrain the operating conditions for the industrial application of this
process. This will open the way to the discussion on the operating conditions

R. Baciocchi (*) • G. Costa • D. Zingaretti


Laboratory of Environmental Engineering, Department of Civil Engineering
and Computer Science Engineering, University of Rome “Tor Vergata”,
Via del Politecnico 1, 00133 Rome, Italy
e-mail: baciocchi@ing.uniroma2.it

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 263
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_11,
© Springer-Verlag Berlin Heidelberg 2014
264 R. Baciocchi et al.

needed to accelerate the process, making it suitable for industrial applications.


The application of accelerated carbonation will then constitute the main core of
this chapter. First, a few CO2 capture processes based on the exploitation of the
carbonation reaction will be discussed. In the following, the largest section will be
devoted to the application of mineral carbonation as a CO2 storage and/or utilisation
process, providing details on the different materials that can be used (minerals or
industrial residues) and the reaction routes applied so far. Finally, the performance
of accelerated carbonation processes making use of specific routes and feedstocks
will be compared in terms of material and energy requirements.

11.2 Fundamentals of Carbonation Processes

11.2.1 Basic Thermodynamics and Kinetics of Carbonation

Mineral carbonation is a naturally occurring process, by which carbon dioxide reacts


with minerals typically found in mafic and ultramafic rocks (i.e. Mg2SiO4 olivine,
Mg3Si2O5(OH)4 serpentine, CaSiO3 wollastonite), forming geologically and thermo-
dynamically stable mineral carbonates (i.e. MgCO3 magnesite, MgCa(CO3)2
dolomite, CaCO3 calcite, FeCO3 siderite, NaAl(CO3)(OH)2 dawsonite) [1]. Carbon-
ation is a thermodynamically favourable process at the moderate temperature and
pressure conditions encountered on the earth’s surface and in the more superficial
earth crust. The free energy of calcium and magnesium minerals, such as wollas-
tonite and forsterite, in fact, well exceeds that of the corresponding carbonates for
temperatures up to about 500 K at 1 bar CO2 partial pressure [2]. Besides, the
carbonation reaction is also exothermic: for instance, the enthalpy of reaction of
CaO to CaCO3 is 90 kJ/mol, which compares well to the enthalpy of combustion
of carbon to carbon dioxide, equal to 400 kJ/mol.
Nevertheless, the naturally occurring carbonation reaction is a slow process,
taking place at geological time scales. According to an investigation by Haug
et al. [3], the weathering rate of olivine is estimated to be 108.5 mol/(m2s)
considering the average ground temperature in Norway of 6  C and a pH of 5.6,
which corresponds to the acidity of rainwater.

11.2.2 Accelerated Carbonation

The acceleration of natural carbonation processes in order to make them suitable for
industrial applications aimed to store anthropogenically derived carbon dioxide has
been first proposed by Seifritz [4] and Lackner et al. [2] within the efforts to reduce
the emissions of greenhouse gases. The accelerated carbonation process involves
the reaction of carbon dioxide with oxides of alkali and alkaline earth metals:

MeO þ CO2 $ MeCO3 ð11:1Þ


11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 265

Usually alkali or alkaline earth metals are not found naturally in the form of free
oxides, as the latter are generally produced by the energy intensive calcination of
the corresponding naturally occurring carbonate minerals. This means that reaction
(11.1) as such can be exploited only if the carbonate is regenerated back to its oxide
form. This is typically the case of CO2 capture processes based on carbonation and
calcination cycles. In these applications, CaO or other free alkali or alkaline metal
oxides are usually contacted with a diluted CO2 source at a temperature high
enough to rapidly convert the oxide to the corresponding carbonate form. Then,
the carbonated material is conveyed to a calciner where it is heated in order to
decompose the carbonate, allowing to regenerate the solid to the oxide form, while
releasing carbon dioxide ready for storage [5]. This process will be discussed in
more detail in Sect. 11.3.2, together with another capture process based on the use
of hot potassium carbonate [6].
If carbonate regeneration cannot take place or if the goal of the process is CO2
storage (see Sect. 11.4.1), another alkalinity source needs to be used. As discussed
in the previous section, alkali and alkaline earth metals may represent such a source.
These are abundant components of the earth’s crust, where they are found in
different minerals, such as serpentine, olivine and wollastonite. The carbonation
of these mineral takes place through the following global reactions [7]:

Olivine Mg2 SiO4 þ 2CO2 $ 2MgCO3 þ SiO2 ð11:2aÞ


Serpentine Mg3 Si2 O5 ðOHÞ4 þ 3CO2 $ 3MgCO3 þ 2SiO2 þ H2 O ð11:2bÞ
Wollastonite CaSiO3 þ CO2 $ CaCO3 þ SiO2 ð11:2cÞ

Although a worldwide comprehensive estimate of available alkaline minerals is


missing, the studies that have been performed up to now suggest that these should
be theoretically sufficient to store all the currently known fossil carbon reserves.
Making reference to the US situation, serpentine reserves are estimated at about
9,700 Gt [8]. Assuming that 2.1 tons of mineral are needed per ton of CO2 stored
(100 % conversion of serpentine), the sequestration potential of this mass of
material is estimated around 4,600 Gt of CO2 or more than 500 years of the US
current production of CO2 [8]. Furthermore, the final products of mineral carbon-
ation (carbonates and silica) are environmentally stable and can therefore be
disposed of as mine filler materials or used for construction purposes [9]. Hence,
unlike underground geological storage, there would be little or no need to monitor
the disposal sites and the environmental risks would be very low [7]. In Fig. 11.1 the
estimated capacities and storage timeframes associated to different sequestration
techniques are represented; as can be noted, mineral carbonation in particular shows
the largest potential both in terms of capacity (similar to underground injection in
saline aquifers) and of storage timeframes.
Despite the large availability of these minerals, concerns have been raised on the
costs and environmental impact of the large mining industry that should be devel-
oped to extract them from the earth’s crust. The size of this industry would not be
unfeasible but would probably be of the same scale of the coal mining industry [7].
266 R. Baciocchi et al.

Fig. 11.1 Estimated storage capacities and relative timeframes for various sequestration methods
(Reprinted from Ref. [10]. Copyright 2006, with permission from Elsevier)

An alternative alkalinity source is represented by some kinds of residues pro-


duced by different industrial activities. These include slags produced by iron and
steelmaking plants (SS), fly and bottom ash (FA and BA, respectively) produced by
power plants and solid waste incinerators and finally by cement kiln dust (CKD) or
even waste cement [11]. Globally, a few hundred Mt of these materials is produced
worldwide and part of it is already reused or recycled. Recently, Kirchofer
et al. [12] provided an upper estimate of total Ca and Mg alkalinity available
in the United States from CKD, FA and SS equal approximately to 7.65 Mt/year,
with a carbonation capacity of approximately 6.98 Mt-CO2/year, compared to about
6 Gt CO2 emitted yearly in the United States. Figure 11.2 also indicates that the
alkalinity and CO2 emissions sources are well overlapped, suggesting that it is
unlikely that carbonation reactants would need to be transported at great distances
[12]. Clearly, these figures suggest that carbonation of industrial residues cannot be
considered as a relevant global option for CO2 storage. Nevertheless, it may help in
reducing CO2 emissions of some specific and small-scale emission sources.
Besides, carbonation of alkaline industrial residues may be considered as a CO2
utilisation option aimed to improve the environmental properties of these materials
and eventually to produce artificial aggregates that could replace natural ones
in several civil engineering applications [13]. Based again on the US market data,
the potential total synthetic aggregate production using FA, CKD and SS is
estimated to be 20 Mt/year, approximately 1.7 % of total US mined aggregate
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 267

Fig. 11.2 Industrial alkalinity sources and stationary CO2 sources in the conterminous United
States (Reprinted with permission from Ref. [12]. Copyright 2013 American Chemical Society)

Fig. 11.3 Ratio between potential synthetic aggregate production from mineral carbonation to
mined aggregate production (i.e. the potential market share of synthetic aggregate) in the conter-
minous United States (Reprinted with permission from Ref. [12]. Copyright 2013 American
Chemical Society)

(i.e. 1.2 Gt/year) [12]. Figure 11.3 shows anyhow that this ratio is an average of a
clearly dishomogeneous situation, characterised by US states where it may be
expected that even more than 10 % of natural aggregates may be replaced by
carbonated materials.
268 R. Baciocchi et al.

11.3 CO2 Capture

11.3.1 Carbon Dioxide Sources and Capture Options

Carbon capture consists in the production of a concentrated CO2 stream from a


diluted source with the aim of conveying it to storage [7]. Diluted sources may
include flue gases from combustion plants with CO2 concentrations ranging
between 3 and 10 % vol. depending on the fuel used and the excess air applied,
flue gases from cement kilns with CO2 concentrations up to 15 % vol. and syngas
from integrated gasification combined cycle (IGCC) plants with CO2 concentra-
tions up to 40 %. The different CO2 concentrations of the various emission sources
pose different challenges in the development and choice of the most suitable
separation process to use. Currently, capture processes based on chemical absorp-
tion using amine solutions represent the benchmark technology as they allow for
the lowest energy penalty with respect to all other alternatives ready for use [7].
Nevertheless, the energy penalty of this process, mainly related to the heat require-
ments needed for amine regeneration, is approximately 25 % of the power plant
thermal duty [7]. For this reason, currently, many efforts are being made to develop
alternative capture processes characterised by lower energy penalties.

11.3.2 Capture Processes Based on Carbonation

This section will discuss three capture processes based on carbonation chemistry:
calcium chemical looping, the Benfield process, based on the use of hot potassium
carbonate, and absorption with alkali solutions. The former one is still under
development; the second one is already applied at full scale, although its application
seems limited to concentrated CO2 sources; the latter one has been proposed for
direct capture from air, although its feasibility is questionable, as will be discussed
in this section.

11.3.2.1 Calcium Looping

Capture using calcium looping has been investigated for more than a decade. This
process is based on two steps. First, carbon dioxide is captured in a carbonation
reactor, where the following reaction takes place:

CaO þ CO2 $ CaCO3 ð11:3Þ

Then, the carbonates are calcined at higher temperature, leading on the one hand
to calcium oxide regeneration and on the other hand to produce a carbon dioxide
stream suitable for disposal:

CaCO3 $ CaO þ CO2 ð11:4Þ


11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 269

Fig. 11.4 Scheme of the calcium looping process applied to precombustion capture (Reprinted
from Ref. [15]. Copyright 2010, with permission from Elsevier)

In the first applications that were not aimed at carbon capture and storage (CCS),
the production of rather pure CO2 was not an issue and the heat required for the
calcination step was supplied by fuel combustion with air [5]. Shimizu et al. [14]
first proposed to couple reactions (11.3) and (11.4) in the framework of CCS,
following the process scheme outlined in Fig. 11.4. Carbon dioxide is captured in
the carbonator at about 600  C and the corresponding carbonates are calcined at
950  C providing the heat required by oxy-fuel combustion, thus leading to a CaO
stream that is recycled to the carbonator and a CO2 stream of about 95 % purity,
excluding the water that can be removed by condensation that can be disposed of.
The application of the calcium looping process requires that the sorbent is
subjected to many cycles of carbonation and calcination [15]. Unfortunately, the
CO2 carrying capacity of the sorbent was shown to decline for increasing number of
carbonation/calcination cycles (see Fig. 11.5). Increasing the regeneration temper-
ature was also shown to negatively affect the performance of the sorbent after
several cycles. This suggests that most of the observed loss of the CO2 carrying
capacity is due to sintering phenomena occurring during the calcination step.
Sintering here refers to changes in pore shape, pore shrinkage and grain growth
that particles of CaO undergo during heating. Sintering of CaO increases at higher
temperatures and durations of calcination, as well as at higher partial pressures of
steam and carbon dioxide [15].
Different options have been proposed to increase the lifetime of calcium-based
sorbents. Among these, hydration of calcium oxide may represent one of the most
promising options. Zeman [16] proposed to include a third hydration step in the
calcium looping process, as reported in Fig. 11.6. Extrapolating the data collected in
270 R. Baciocchi et al.

Fig. 11.5 CO2 carrying


capacity of the sorbent at
increasing number of
carbonation/calcination
cycles (Reprinted from Ref.
[15]. Copyright 2010, with
permission from Elsevier)

Fig. 11.6 Scheme of the


hydrated calcium looping
process (Reprinted from
Ref. [16]. Copyright 2008,
with permission from
Elsevier)

ten TGA carbonation/calcination/hydration cycles, it was suggested that the


long-term sorbent residual carrying capacity could increase up to 600 % with respect
to dry carbonation/calcination cycles. According to Blamey et al. [15] this could in
principle allow to reduce the sorbent makeup from the 1 mol of sorbent capturing
4.33 mol of CO2 over ten cycles reported by Abanades et al. [17] up to 7.20 mol of
CO2 over ten cycles when the hydration step is included. Another interesting possi-
bility to increase the lifetime of the sorbent is to develop synthetic products. One
option relies on the use of pure calcium carbonate (PCC) that will also be discussed
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 271

later in the framework of CO2 utilisation for the carbonation of industrial residues.
High surface area PCC particles (5–20 nm) were synthetised through a wet precip-
itation process, allowing to achieve over 90 % carbonation over two calcium looping
cycles [18]. The lifetime of the sorbent can also be improved by using mixed oxides
with the formula Ca0.9 M0.1Ox where M ¼ Cu, Cr, Co or Mn, CaO sorbents derived
from organic salt precursors, CaO doped with Cs, sorbents derived from calcium
lignosulphonate, sorbents based on Na2CO3 or NaCl, CaO-based sorbents using
CaTiO3 as binder and CaO-based sorbents using calcium aluminate (Ca12Al14O33)
as a stable matrix (see [19] and references therein).

11.3.2.2 Absorption with Hot Carbonate Solutions

The capture process is, in this case, based on the reaction of carbon dioxide with a
potassium carbonate solution [6]:

CO2 þ K2 CO3 þ H2 O $ 2KHCO3 ð11:5Þ

The CO2 absorber is usually operated at high temperature and pressure. The
higher the temperature is, the higher the concentration of potassium carbonate that
can be used without precipitating the carbonate [20]. As the driving force of the
process is given by the difference between the CO2 partial pressure in the gas bulk
and its equilibrium pressure at the gas–liquid interface, the process is usually
operated at high pressure. Typical operating conditions of the absorber are
120–125  C and 25 atm [20]. The bicarbonate solution is regenerated by steam
stripping at approximately the same operating temperature and recycled to the
absorption unit, whereas the CO2 liberated may be disposed of. The energy require-
ment needed to regenerate thermally the bicarbonate back to carbonate is around
3–4 GJ/t CO2 [21].
The chemistry of the hot carbonate absorption process is clearly in aqueous
phase. At the typical pH (>8) encountered in the absorbing solution, the following
reaction represents the rate-limiting step:

CO2 þ OH $ HCO


3 ð11:6Þ

The CO2 mass transfer rates may be improved through additives to the potassium
carbonate solution [20]. These may include small amounts of diethanolamine
(DEA), which significantly enhances the absorption rate, through the following
reactions, leading to the formation of carbamate that then reacts with hydroxide
ions forming back DEA and leading to the formation of bicarbonate ions:

CO2 þ R2 NH $ R2 Nþ HCOO ð11:6aÞ


þ  
R2 N HCOO þ OH $ R2 NH þ HCO
3 ð11:6bÞ

The Benfield industrial process, based on the chemistry discussed above, has
been so far applied in 700 industrial plants for natural gas sweetening and in
272 R. Baciocchi et al.

ammonia plants. The relatively high partial pressure required for maximising the
effectiveness of this process makes it more suitable for processing concentrated gas
streams. For this reason, the Benfield process is used to treat the gas from the water-
gas shift converter downstream a steam-reforming unit, where the CO2 must be
removed in order to obtain a pure nitrogen/hydrogen gas mixture for ammonia
synthesis. In the framework of CCS technologies, it could be used as well to capture
the CO2 from syngas obtained from coal gasification units.

11.3.2.3 Absorption with Alkali Solutions

Carbon dioxide can be captured also using alkali solutions, such as NaOH or KOH
ones. In this case, carbon dioxide is neutralised through the following general
reaction:

CO2 þ NaOH $ Na2 CO3 þ H2 O ð11:7Þ

As already discussed for CaO, also NaOH is produced through an energy


intensive process. Therefore, the regeneration of the carbonate to the hydroxide
form is needed to make the process sustainable in terms of energy requirements and
carbon capture. The traditional approach for the regeneration of potassium carbon-
ate comes from the paper industry and is based on the so-called causticisation
reaction, where potassium carbonate is reacted with lime:

Na2 CO3 þ CaðOHÞ2 $ NaOH þ CaCO3 ð11:8Þ

This reaction allows to regenerate NaOH, but once again requires lime as
reagent, that is produced from hydration of quicklime (CaO). The latter, as already
discussed, is produced by calcination of calcium carbonate (see Eq. 11.4).
The capture process given by Eqs. (11.7), (11.8) and (11.4) has been proposed to
capture carbon dioxide from air and has been first assessed by Baciocchi et al. [22]
and more recently has been the subject of a study of the American Physical
Society [23]. The scheme of the process used as reference in both cited works is
reported in Fig. 11.7.
In principle, the application of the alkali absorption process is well suited for
CO2 storage from very diluted sources, as air where CO2 is present at about
380–390 ppm. This stems from the fact that hydroxyl ions accelerate the rate of
CO2 transfer from the gas to the liquid phase [24], which would be very slow if
based on physical absorption or chemical absorption using weak bases, such as
ethanolamines. Nevertheless, the main limitation of this capture process relies in
the very high energy needed for regeneration. Baciocchi et al. [22] showed that this
should amount to 12 GJ/t CO2 against 9 and 20 GJ/t CO2 generated by a power plant
operating with coal or natural gas, respectively, and much higher than the 3–4 GJ/t
CO2 reported for flue gas capture using amines or hot potassium carbonate [21],
thus making the process absolutely unfeasible in the first case and characterised by
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 273

H2O make-up

1B 1C

2-slaker
1-absorber

2B
2A 6B 8-condenser
heat
exchanger 7F
3A 7E
1A (air) 1D 3B
3C
splitter 8B
7-kiln
8A
6B 7D
3-pellet reactor 9A

7C (solid makeup)

7A+7B (fuel + oxygen from PSA) 9-compressor

Fig. 11.7 Scheme of the air capture process based on alkali absorption and regeneration by
causticisation and calcination reactions (Adapted and reprinted from Ref. [22]. Copyright 2006,
with permission from Elsevier)

a too high energy penalty in the latter one. This conclusion was reinforced by APS
[23] that gave a realistic estimate of air capture using alkali solutions of around
780 USD/t CO2. In order to become a feasible capture option, the alkali absorption
process would require a less energy intensive regeneration step. Nagasawa
et al. [21] investigated how to increase its efficiency using a regeneration process
based on the acidification of the carbonate solution obtained from the absorber,
followed by liberation of gaseous carbon dioxide.

CO2 þ Hþ ! HCO
3 ð11:9aÞ
HCO
3
þ
þ H ! H2 CO3 ð11:9bÞ
H2 CO3 ! H2 O þ CO2 ð11:9cÞ

As the reaction requires protons on the one hand and to remove alkaline metal
ions from the solution on the other hand, this can be accomplished using electro-
dialysis for such an exchange. After an optimisation of the cell configuration,
Nagasawa et al. [21] reduced the energy requirements to around 4 GJ/t CO2.
Nevertheless, the feasibility and applicability of this process need yet to be
demonstrated.
274 R. Baciocchi et al.

11.4 CO2 Storage

11.4.1 Processes and Reaction Routes

Accelerated carbonation for carbon storage has been investigated using different
reaction routes (see Fig. 11.8). These can be broadly classified according to the
reacting phases in gas-solid and aqueous phase routes. In the former one, gaseous
carbon dioxide is assumed to react directly or indirectly with an alkaline mineral or
residue. This route can be applied to materials characterised by a relatively high
content of free calcium oxides or hydroxides. For the other alkaline materials, it is
possible to apply the indirect route, where the reaction goes through the conversion
of the alkaline material to the corresponding oxide or hydroxide, which is then
carbonated in a subsequent step. As it will be discussed in more detail in the next
sections, the gas-solid route does not seem promising for the scaleup of the process,
in view of the slow kinetics of carbonation observed for both the indirect and direct
options. As a consequence, most of the research on accelerated carbonation has

Fig. 11.8 Main routes of the accelerated carbonation processes tested for CO2 storage
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 275

focused on the aqueous route, generally mixing the alkaline feedstock with water
applying liquid to solid (L/S) ratios above 2 l/kg (slurry phase). Also in this case,
both direct and indirect reaction routes have been proposed and tested. In the first
case, mineral dissolution and carbonate precipitation take place in the same reactor.
Eventually, chemicals, such as salts and bicarbonates, may be added in order to
enhance the rate of the reaction by buffering the solution pH. As the optimal
operating conditions for dissolution and precipitation are quite different, the indi-
rect route may be preferred, as it allows running the dissolution and precipitation
reactions in two different steps, switching the operating conditions as required.
Finally, it is worth noting that the aqueous direct route may be run also using very
low L/S ratios (typically below 1 l/kg), and in this case, it is often referred to as the
“wet route”. This route has been particularly investigated for the carbonation of
industrial residues and will be discussed in Sect. 11.4.4.

11.4.2 Mineral Carbonation

11.4.2.1 Gas–Solid Carbonation

As discussed in Sect. 11.2.1, carbonation of minerals is thermodynamically


favourable even at ambient conditions, with an exothermic overall chemistry that
can be summarised as follows [25]:

xMgO ySiO2 zH2 OðsÞ $ xMgOðsÞ þ ySiO2 ðsÞ þ zH2 O ð11:10aÞ


MgOðsÞ þ CO2 $ MgCO3 ðsÞ ð11:10bÞ

Direct gas-solid carbonation of Mg silicate minerals would have been the


preferred route for mineralisation of CO2 due to its simplicity in design and better
energy efficiency. More so, the overall carbonation reactions of olivine and ser-
pentine are exothermic. However, these reactions suffer from very slow kinetics,
low conversion and high energy requirements [26]. For this reason, Nduagu
et al. [26] proposed an indirect gas-solid process based on the following steps:
(1) production of MgSO4 from the silicate, (2) conversion of MgSO4 to Mg(OH)2
and (3) gas-solid carbonation of Mg(OH)2 to MgCO3. One of the limitations of this
process is given by the heat required to produce Mg hydroxide from the
corresponding silicate, a process that requires temperatures of around
400–500  C. Therefore, most of the feasibility of the process route comes from
the possibility of using the heat released by the exothermic carbonation reaction.
Besides, it is worth noting that the rate of Mg(OH)2 carbonation is quite slow, with
reasonable conversions achieved only after 6 h of reaction at about 35 bar and
500  C (see Table 2 in [25]).
276 R. Baciocchi et al.

11.4.2.2 Aqueous Phase Carbonation

This process route takes place through the following chemical reactions, written
making reference to forsterite, the main mineral constituent of olivine:
(a) Dissolution of carbon dioxide in water with release of protons:

CO2ðgÞ þ 2H2 OðlÞ ! H2 CO3ðaqÞ ! Hþ ðaqÞ þ HCO


3ðaqÞ

(b) Dissolution of the mineral in water with release of alkaline cations:

Mg2 SiO4ðsÞ þ 4Hþ ðaqÞ ! 2Mg2þ ðaqÞ þ SiO2ðsÞ þ 2H2 OðlÞ

(c) Precipitation of the carbonates:

Mg2þ ðaqÞ þ HCO þ


3ðaqÞ ! MgCO3ðsÞ þ H ðaqÞ

In the direct aqueous phase route, the different reactions are assumed to take
place in the same reactor at the same operating conditions. A comprehensive study
of this route can be found in the works of the Albany Research Center [27] and of
the Energy Research Centre of the Netherlands [28–30]. Gerdemann et al. [27]
investigated the carbonation of different mineral feedstocks, including olivine,
serpentine and wollastonite. They performed tests at different pressure, temperature
and L/S ratio, which allowed them to select the best operating conditions in terms
of conversion of the alkaline reactant (Mg or Ca depending on the feedstock).
Namely, these were achieved at T ¼ 185  C and pCO2 ¼ 150 atm for olivine and
T ¼ 155  C and pCO2 ¼ 115 atm for serpentine, using in both cases a 0.64 M
NaHCO3 and 1 M NaCl solution. For wollastonite the optimal operating conditions
were slightly milder (T ¼ 100  C and pCO2 ¼ 40 atm) and distilled water was
used. It is worth pointing out that the best operating conditions for serpentine were
obtained assuming to use heat-activated serpentine (above 630  C) and in all cases
finely ground feedstocks (below 75 μm). Besides, for both olivine and serpentine, a
pH buffer solution was needed to control the pH allowing both dissolution and
carbonation to occur at a reasonable extent. The maximum extents of reaction
achieved were around 50 % for olivine, 80 % for wollastonite and 73 % for
serpentine. Huijgen et al. [28, 29] also investigated the carbonation of wollastonite
in a direct route process. Wollastonite was observed to carbonate rapidly, with a
maximum conversion in 15 min of 70 % at relatively mild conditions (d < 38 μm,
T ¼ 200  C and pCO2 ¼ 20 bar).
The works performed on the direct route made clear that the limiting step of the
whole process is represented by the dissolution of the mineral, especially for
Mg-bearing ones. An increase of the dissolution rate can be achieved by performing
a physical/chemical mineral surface pretreatment, increasing the specific area (grind-
ing for size reduction), or by removing the silica-passivating layer (by abrasion/
attrition). The first three options allow basically to modify the physicochemical
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 277

Fig. 11.9 Dissolution rate of olivine, r [mol cm2 s1], vs. pH at different temperatures. Activation
energy: 52.9 kj/mol (Adapted and reprinted from Ref. [31]. Copyright 2006, with permission
from Elsevier)

properties of the material surface prior to dissolution. The latter one may allow to
remove the residual silica layer on the mineral surface that prevents further dissolu-
tion of magnesium. For this reason, several works have been performed in order to
better understand the dissolution mechanism and to increase the dissolution rate,
especially of Mg-bearing minerals. Hänchen et al. [31] have specifically investigated
the dissolution of olivine, showing the effect of the solution pH on the rate of
dissolution at operating conditions suitable for mineral carbonation (see Fig. 11.9).
Besides, the same group has more recently demonstrated that the addition of salts to
the solution does not help in improving the dissolution rate of olivine [32], differently
from what was expected based on the findings of Gerdemann et al. [27].
The findings on olivine dissolution are a clear example of the fact that the rate of
silicate minerals dissolution is clearly enhanced at lower pH values, where precip-
itation of carbonates can hardly take place. This issue has prompted the develop-
ment of a few indirect carbonation schemes, where minerals dissolution and
carbonate precipitation are decoupled. Park et al. [33] developed a pH swing
process, whose scheme is reported in Fig. 11.10. Dissolution of serpentine is
278 R. Baciocchi et al.

Dissolution
- 1% vol. Ortoposphoric acid Product 1
- 0.9% wt Oxalic acid Undissolved solids
- 0.1% wt. EDTA (rich in SiO2 content)
- internal grinding (2mm glass
beads) –T= 70°C
NH4OH CO2 (1 atm)
Mg-and Fe-rich solution

NH4OH Precipitation Precipitation


pH=8.6 pH=9.5

Product 2 Product 3

Iron oxide MgCO3*3H2O

Fig. 11.10 Scheme of the pH swing process for serpentine carbonation developed by [33]

enhanced using a solution of weak acids and chelating agents coupled with internal
grinding using glass beads. Then, two precipitation steps are foreseen: the first one,
at pH ¼ 8.6, aimed at iron oxide precipitation and the second one, at a higher pH,
aimed at nesquehonite precipitation. Apart from the complexity of the different
steps, the main open issue of this process remains how to recycle the different acid
and basic amendments needed.
This is also the case for the other indirect process routes proposed in the last
decade. Teir et al. [34] dissolved serpentine in HCl or HNO3, and then
hydromagnesite (4MgCO3 Mg(OH)2 4H2O) was obtained by fixing the Mg-rich
solution pH to 9 with addition of NaOH. Although a high carbonation conversion
efficiency of Mg (80–90 %) was obtained, the chemicals involved, such as HCl and
NaOH, could not be reused. More recently, Wang and Maroto-Valer [35] proposed
a new pH swing mineral process by using recyclable ammonium salts. The process
consists of the following main steps: (a) CO2 is captured from flue gas using NH3
and producing NH4HCO3; (b) a solution of NH4HSO4 is used to extract Mg from
serpentine at mild operating conditions leading to a MgSO4 solution; (c) the pH of
the Mg-rich solution produced from mineral dissolution is switched to neutral pH
by adding NH4OH; (d) the MgSO4 solution reacts with the NH4HCO3 solution
allowing to precipitate carbonates at mild temperature, in fact since the formation
of carbonate is determined by the operating temperature [36], nesquehonite
(MgCO3·3H2O) can be converted to hydromagnesite (4MgCO3·Mg(OH)2·4H2O)
above 70  C; and (e) the (NH4)2SO4 remaining in solution after the precipitation of
hydromagnesite could be recovered by evaporation and subsequently heated up to
regenerate NH3 which is fed back to the capture step, whereas NH4HSO4 could be
then reused for the mineral dissolution step.
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 279

11.4.3 Carbonation of Industrial Residues

As mentioned in Sect. 11.2.2, an alternative feedstock of alkaline phases suitable to


be treated by accelerated carbonation for CO2 storage is represented by waste
by-products characterised by high calcium or magnesium (hydr)oxide or silicate
contents; these include residues from thermal treatment plants or other activities
such as mining, metallurgical processes, cement or paper manufacturing and
construction and demolition [11, 37, 38]. Besides being often associated with
CO2 point source emissions, these materials tend to be chemically more unstable
than geologically derived minerals [28]; therefore, typically they require a lower
degree of pretreatment and less energy intensive operating conditions to enhance
carbonation yields [28, 39]. Although as previously mentioned, the availability of
these types of residues is too limited to exert a substantial effect in terms of the
reduction of global CO2 emissions, their use as feedstock for accelerated carbon-
ation processes could be seen as a stepping stone towards the development of CCS
processes based on natural minerals [37]. Furthermore, several studies have shown
that accelerated carbonation is an effective treatment for reducing the release of
specific metals from alkaline waste materials such as waste incineration bottom ash
[40–43] and Air Pollution Control (APC) residues [44–46], steel slag [47–50] and
oil shale ash [51], among others. The improvement in the leaching behaviour of
these waste materials exerted by accelerated carbonation is important since it may
broaden their management options, allowing e.g. to reuse them in selected civil
engineering applications or to dispose of them in landfills for non hazardous or even
inert waste. In addition, following specific reaction routes, accelerated carbonation
can lead to the formation of products that can be employed for different applica-
tions, such as aggregates in construction or calcium carbonate to be used as filler or
coating pigment in paper or plastics manufacturing.

11.4.3.1 Tested Materials

Accelerated carbonation has been tested on a variety of residues as a technique


to achieve different objectives in different countries. In fact, issues of primary
importance for this technology are the relative abundance of each waste stream,
which determines the overall potential impact of the accelerated carbonation
technique as a CO2 mitigation strategy and the environmental behaviour of the
residues. Both these issues may vary from country to country as a function of many
factors including the main features of the industrial and energy sectors, available
natural resources, implemented waste management strategies, adopted waste resi-
due disposal and flue gas treatment technologies as well as environmental legisla-
tion. Thus, for example, in Estonia most of the energy demand is supplied by
oil shale combustion, which emits significant amounts of CO2 (emission factor of
29.1 tC/TJ) and generates large amounts of alkaline oil shale ash, thus making CO2
sequestration via carbonation of oil shale ash a very attractive process [52];
280 R. Baciocchi et al.

Fig. 11.11 Normalised CaO/MgO–SiO2–Al2O3/Fe2O3 phase diagram of various types of alkaline


waste (Reprinted with permission from Ref. [38]. Copyright 2012, Taiwan Association for Aerosol
Research)

in Belgium instead, stringent legislative landfill acceptance criteria related to the


leaching of specific heavy metals (such as Cu, Zn and Pb, in particular) do not allow
waste incineration bottom ash to be directly landfilled even after natural weathering
periods; hence, accelerated carbonation studies have been specifically undertaken
to improve the chemical stability of these residues [40, 43, 53].
One of the fundamental aspects of accelerated carbonation investigations on
alkaline industrial waste materials is the physical, chemical and mineralogical
characterisation of the residues for the identification and quantification of the
reactive phases. As shown in Fig. 11.11, the residues that exhibit the highest
contents of CaO and MgO are municipal solid waste incineration (MSWI) fly ash
or APC residues, basic oxygen furnace (BOF) steel slag and cement kiln dust
(CKD) that present quite a similar content of major constituents to ordinary
Portland Cement (PC). Other types of metallurgical residues such as
desulphurisation (De-S) and blast furnace (BF) slags also present significant CaO
and MgO contents, whereas MSWI bottom ash and coal combustion fly ash
generally exhibit a predominance of Si phases and significant contents of Fe and
Al oxides [38]. Some types of solid residues streams produced by coal-fired power
stations, however, such as lignite combustion fly ash and oil shale ash typically
exhibit higher CaO and MgO contents, i.e. 37 % CaO and 15 % MgO the
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 281

50

45

40

35
CO2 uptake (% wt.)

AOD
30
BOF CBD
25 APC
BF LF
20
EAF
15
CKD
10 MSW-FA RDF-BA
CC-FA
5 MSWI-BA
RM
0
0 10 20 30 40 50 60 70
CaO content (% wt.)

Fig. 11.12 Comparison of the average CO2 uptakes obtained for different types of residues as a
function of the CaO content of the materials, in which the purple-dotted line reports the maximum
CO2 uptake expected on the basis of the CaO content of the residues. Materials: CKD cement kiln
dust [11], CBD cement bypass dust [11], APC air pollution control residues [45], RDF-BA refuse-
derived fuel bottom ash [42], MSWI-BA municipal solid waste incineration bottom ash [11],
MSWI-FA municipal solid waste incineration fly ash [11], RM red mud [38], BOF basic oxygen
furnace slags [38], EAF electric arc furnace slag [48], AOD argon oxygen decarburisation slag
[48], BF blast furnace slag [38], LF ladle furnace slags [38], CC-FA coal combustion fly ash [38]

former [54] and around 50 % CaO and 10–15 % MgO the latter type of material
[52]. Accelerated carbonation has been applied also to alumina manufacturing
residues, generally known as red mud, which typically contain around 8 % CaO
and 1 % MgO [55].
The main composition of a residue can already provide a general indication on
its reactivity with CO2. As can be noted in Fig. 11.12 in fact, the average CO2
uptakes resulting for each type of tested residue can be directly related in most cases
to their CaO content, since Ca-based phases generally present a higher reactivity
even under rather mild operating conditions, as opposed to Mg, Fe and Al ones.
Hence, the residues that have shown to yield the highest CO2 sequestration potential
are those generated in steel-refining processes, such as AOD and LF slag, APC
residues and by-products from cement manufacturing such as CBD (see Fig. 11.12).
However, besides bulk chemical composition, important factors affecting the
reactivity of industrial residues with CO2 are also their mineralogy and particle
size distribution. As previously mentioned, free oxide and hydroxide phases such
as lime and portlandite, typically found in APC residues, CKD and paper mill
waste, are very reactive with CO2 even at mild operating conditions, while the
reactivity of many of the different Ca- and Mg-containing silicates detected in
282 R. Baciocchi et al.

metallurgical slag, construction and demolition waste and MSWI bottom ash vary
greatly depending on the types of crystal phases and of the presence of inclusions in
the crystal lattice of various elements, such as Al, Cr and Fe. The reactivity of Ca
silicates is indicated to increase for higher Ca/Si ratios [56]; in fact the most
reactive silicate phases with CO2 are generally reported to be tricalcium (alite,
Ca3SiO5) and dicalcium (belite, Ca2SiO4) silicates, the latter in particular in the
γ-polymorph phase that is formed during the cooling of steel-refining slag and
results in the disintegration of the slag into a fine powder [48]. These types of
silicate phases are generally retrieved in materials containing cementitious phases,
such as concrete demolition waste [57] or steel manufacturing slag. Particle size, as
indicated in Sect. 11.4.2, is one of the controlling factors for the dissolution kinetics
of any kind of mineral or anthropogenically derived material. The by-products of
flue gas treatment units, cement kiln dust and other ashes or particular products of
industrial refining processes present average grain sizes (generally <100–150 μm)
that are already in the optimum range for carbonation treatment, whereas slag
residues from combustion or metallurgical processes or scraps from construction
and demolition activities present a wider particle size distribution curve and a
significant percentage of coarse particles.

11.4.3.2 Investigated Reaction Routes

Accelerated carbonation of alkaline solid waste has been up to now primarily


investigated through the direct aqueous route (e.g. [40, 44, 45, 48, 52, 54,
58, 59]) or to a lower extent by means of the direct gas-solid reaction, applied
only to materials with high lime and portlandite contents [60–62]. As mentioned in
Sect. 11.4.1, direct aqueous accelerated carbonation studies have been performed in
two different modes, either in slurry phase applying liquid to solid (L/S) ratios
above 2 w/w (e.g. [54, 55, 58, 63, 64]), in particular for waste materials of low
solubility in which CaO is generally bound as a silicate, or via the wet route,
humidifying the material so as to set the L/S ratio below 1 w/w (e.g. [41, 42, 44,
45, 48, 49, 51, 59]). This latter type of treatment in which the CO2 and Ca and Mg
ions dissolution, as well as carbonation reactions, occur in the thin liquid film in
direct contact with the solid residues has been originally applied for cement curing
processes (see e.g. [65, 66]). It is specifically applied as a carbonation route also for
industrial residues that present high contents of soluble elements such as salts and
heavy metals so as to avoid the treatment and disposal of the processing liquid,
as well as to favour dissolution kinetics at relatively mild operating conditions.
Carbonation of residues does not generally require the extraction of alkaline
metals from the solid matrix, since the main reactive species are Ca-containing
silicates, oxides and hydroxides. These minerals, as earlier mentioned, behave
differently from those containing Mg, as hydration, solvation and carbonation of
Ca-bearing phases can be carried out under the same operational conditions in a one
stage operation [67]. However, for specific waste residues, characterised by high
contents of silicate phases such as various types of steel slag or cement waste
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 283

residues, accelerated carbonation has also been tested applying the indirect aqueous
route to slurry suspensions, separating the dissolution and precipitation steps so as
to optimise each set of reactions by operating on the pH, dosing specific additives,
similarly to the previously described pH swing process, (e.g. [68, 69]) or by varying
some of the operating parameters, such as CO2 pressure [57].
Depending chiefly on the main composition of the residues, the chosen process
route and also the final aim of the study (CO2 sequestration or improvement of
the properties of the residue), a variety of operational conditions have been applied
and quite different results in terms of CO2 uptake have been achieved (see
e.g. [37, 38, 67]).

Gas–Solid Carbonation

As previously mentioned, direct gas–solid carbonation was tested only on residues


characterised by a high content of Ca (hydr)oxide phases and fine particle size.
Jia and Anthony [60] carried out experiments with coal and petroleum coke
fluidised bed combustion (FBC) fly ash in a pressurised thermogravimetric analyzer
(PTGA) at 1 or 11 bar of 100 % CO2; hydrated ash was found to react faster with
CO2 than unhydrated ash, presenting also a higher final CaO to CaCO3 conversion
ratio (60 % compared to 27.4 % respectively), owing to the effects of hydration on
increasing the ash surface area [60]. Carbonation reactions showed to occur in the
300–600  C temperature range for hydrated ash and in the 400–700  C range for the
unhydrated ash, while CO2 pressure did not show to exert a significant effect on
carbonation kinetics or the final calcium conversion yield [60]. More recently,
Baciocchi et al. [61] and Prigiobbe et al. [62] performed gas-solid carbonation
tests employing APC residues and confirmed that, under these conditions, carbon-
ation is effective at temperatures equal to or above 350  C. Maximum conversions
between 60 and 80 % were obtained at temperatures between 350 and 500  C and
ambient pressure also applying diluted gas streams, i.e. characterised by a CO2
concentration of 10–50 vol.% [62]. In addition, the kinetics of the reaction proved
to be fast, achieving completion in less than 5 min and exhibiting a rapid chemically
controlled reaction followed by a slower product layer diffusion-controlled process
(see Fig. 11.13).

Slurry-Phase Carbonation

Slurry-phase accelerated carbonation has been performed mainly either under


enhanced operating conditions (i.e. T ¼ 50–225  C and pCO2 ¼ 5–30 bar)
(e.g. [43, 50, 58]) or alternatively at ambient temperature and atmospheric pressure
using 10–100 % vol. CO2 streams (e.g. [46, 52, 55]).
Slurry-phase carbonation experiments on ground BOF slag under enhanced
conditions were performed in an autoclave reactor to assess the maximum CO2
uptake of the material, as well as the corresponding Ca conversion yield and the
284 R. Baciocchi et al.

Fig. 11.13 Calcium conversion as a function of time for 10 % CO2 flows at different temperature:
350  C (- - -); 400  C (—); 450  C (···) and 500  C (-·-·-) (Reprinted from Ref. [62]. Copyright
2009, with permission from Elsevier)

influence of operating conditions on process kinetics [58]. A maximum CO2 uptake


of about 180 g CO2/kg BOF slag, corresponding to a carbonation conversion of
74 % of the Ca content of the slag, was achieved in 30 min at 19 bar CO2 pressure
and 100  C for a slag particle size <38 μm. The reactive phases in the slag were
found to be portlandite and Ca-(Fe) silicates and the only reaction product detected
by XRD and SEM/EDX analysis was calcite [58]. As shown in Fig. 11.14, slag
particle size was found to exert the most significant effects on the reaction, smaller
grain sizes proving more reactive towards CO2, while temperature, which enhanced
silicate dissolution up to 200  C, the L/S ratio and CO2 partial pressure showed
milder effects on Ca conversion yields. The diffusion of calcium through the
solid matrix towards the surface of the slag particles appeared to be the rate-
limiting reaction step, which was found to be hindered by the formation of a
CaCO3-coating and a Ca-depleted silicate layer on slag particles during the car-
bonation process [58]. A recent study confirmed that several Ca and/or Mg silicate
phases (i.e. dicalcium silicate, bredigite, merwinite and cuspidine) as well as
periclase, typically found in steelmaking slags such as AOD and continuous casting
(CC) residues, present a significant reactivity to slurry-phase carbonation carried
out under enhanced conditions (T ¼ 90  C, pCO2 ¼ 6 bar and L/S ¼ 16 l/kg,
t < 2 h) [50].
The CO2 sequestration potential of metallurgical slags and of other types of
residues such as fly ash from thermal treatment plants has also been investigated via
the slurry-phase route at milder operating conditions. Carbonation of EAF and ladle
slag carried out at 20  C and ambient temperature with 15 % vol. CO2 flows of 5 ml/
min applying an L/S ¼ 10 kg/kg resulted in a CO2 uptake capacity of 17 g/kg EAF
slag in 24 h and 247 g/kg ladle slag in 40 h, which was ascribed to the higher content
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 285

Fig. 11.14 Calcium conversion and CO2 uptake as a function of the L/S ratio, particle size,
reaction temperature (including Arrhenius-plot with A ¼ 6 %Ca/min and Ea ¼ 3.6 kJ/mol) and
carbon dioxide pressure (Reprinted with permission from Ref. [58]. Copyright 2006 American
Chemical Society)

of portlandite of the ladle slag suspension [63]. Studies on the CO2 sequestration
capacity of red mud reported a CO2 uptake of 40 g CO2/kg residues applying an
L/S ¼ 10 kg/kg at 20  C and ambient pressure, bubbling a 15 % vol. CO2 gas
mixture with a flow rate of 5 ml/min for up to 24 h, which was attributed to the
complete carbonation of the Ca and Na (hydr)oxides content of the residues [55].
Slurry-phase carbonation experiments carried out on fly ash from coal combustion
power plants also showed a high reactivity of the lime content of the ash (over
80 %) with pure CO2 at 30  C and atmospheric pressure for 2 h; owing to the limited
amount of CaO in the ash though, a CO2 sequestration capacity of only 26 g CO2/kg
ash was reported [64]. For oil shale ash carbonation, 89–100 % and 48–73 %
extents of carbonation were reported for experiments conducted on circulating
FBC ash and pulverised FA, respectively, introducing a 10 % CO2 gas mixture
simulating flue gas in an aqueous slurry (L/S ¼ 10 l/kg) at atmospheric pressure
and room temperature for 20–40 min [52]. Lime was found to be the main CO2
binding component, but also periclase and Ca silicates (CaSiO3 and Ca2SiO4)
were taken into account for calculating the maximal binding potential of oil shale
ash [52]. The measured CO2 uptakes varied between 100 and 160 g CO2/kg
residues, depending on the type of material tested. For lignite fly ash, a significantly
higher CO2 uptake (230 g CO2/kg ash, with a Ca conversion yield of 75 %) was
measured in a slurry system at 75  C, using 10–30 % CO2 at atmospheric pressure
286 R. Baciocchi et al.

Fig. 11.15 Kinetics of CO2 uptake and Ca conversion as a function of pressure (T ¼ 50  C,


L/S ¼ 0.4 l/kg) for EAF and AOD slag (Reprinted from Ref. [48], with kind permission from
Springer)

for over 4 h [54]; the authors assessed that this result could account for a CO2
sequestration capacity of 3.5 million tons of CO2, corresponding to around 2 % of
the CO2 emissions from lignite power combustion in Germany [54].

Wet-Route Carbonation

Carbonation following the wet route (also called thin-film method [50]) has been
tested on a wide range of residues, generally applying L/S ratios between 0.1 and
1 l/kg and mild operating conditions (30–50  C and 1–10 bar), to investigate
potential CO2 uptakes as well as the effects on the leaching behaviour and on the
mechanical properties of the materials. For stainless steel slag, Johnson et al. [70]
found a maximum CO2 uptake of 180 g/kg residue for milled (<125 μm) slag
applying an L/S of 0.125 l/kg and 3 bar of 100 % CO2 for 1 h, which led to a
significant increase of the 28-day unconfined compressive strength of the material
(around 10 MPa compared to below 0.6 MPa). The carbonation kinetics of different
types of stainless steel slag was studied adopting a similar experimental approach
by Baciocchi et al. [48]. An increase in the reaction temperature (to 50  C) and L/S
ratio (to 0.4 l/kg) showed to increase the reactivity of Ca silicate phases allowing to
achieve higher CO2 uptakes, whereas, as showed in Fig. 11.15, CO2 pressure did
not appear to exert significant effects, except for the AOD slag for reaction times
above 1 h. The maximum Ca conversion yields achieved for the AOD slag appeared
to be considerably higher than those resulting for milled EAF slag (d < 0.150 mm)
and were related to the differences in the mineralogy of the two slags: the AOD
residues mostly made up by γ-dicalcium silicate, while the others containing
various Ca and Mg silicates [48]. Column carbonation experiments on two types
of converter steel slag were also recently reported [49]. These experiments were
carried out at atmospheric pressure, feeding 900 g of wetted steel slag with a water-
saturated CO2/Ar mixture at a flow rate of 400 ml/min for reaction times ranging
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 287

from 8 to 200 h and temperatures between 5 and 90  C. The aim of this particular
study was to assess the leaching behaviour of the slag more than its CO2 storage
potential; hence, relatively large (2–3.3 mm) steel slag grains were used and
consequently lower carbonation degrees were achieved compared to other studies.
CO2 uptakes of 6 or 15 g/kg slag were reported for the two different types of
converter steel slag in dependence of the free lime content of the material. Reaction
kinetics showed to be enhanced by increasing temperature and humidity [49].
The CO2 sequestration capacity of cement kiln dust (CKD) treated with humid-
ified CO2 was recently assessed in batch and column tests at ambient temperature
and pressure applying gas compositions with CO2 contents varying from 14 to 80 %
[59]. The predominant reactive phase was found to be Ca(OH)2 and, similarly to
what was reported in other studies (e.g. [62]), reaction kinetics was found to follow
the unreacted core model theory, exhibiting at early times a first-order trend
indicating kinetics control by the chemical reaction and at longer times a zero-
order trend induced by diffusion control due to coverage of the reacting particle
surface by calcite [59]. A sequestration extent of over 60 % was observed within 8 h
and the authors estimated that applying this process, a 6.5 % emission reduction in
US cement-related emissions could be attained [59]. Compositions similar to those
indicated in the previous study for cement kiln dust (around 50 % Ca(OH)2) were
indicated also for paper mill waste, which allowed to achieve a CO2 net uptake of
218 g CO2/kg paper waste, with pure CO2 at 30  C and atmospheric pressure for 2 h.
A sequestration potential of round 15,000 tons of CO2 for an average size cellulose
pulp factory was estimated [71].
Wet-route carbonation of industrial residues has also been tested at pilot scale
for specific applications. Mostbauer et al. [72] have investigated carbonation of the
fine fraction of MSWI bottom ash as a landfill biogas-upgrading technique, i.e. to
increase its methane content by removing CO2. In addition, the treatment proved
effective in reducing the H2S content of the biogas and in improving the leaching
behaviour of the slag [72]. Reddy et al. [73] tested carbonation in one of the largest
coal-fired power plants of the United States by reacting flue gas with coal fly ash in
a fluidised bed reactor for 1.5–2 h at 60  C. Upon the treatment, the CO2, SO2 and
Hg concentrations of the flue gas showed to decrease [73]. The feasibility of both
types of applications mentioned, however, appears to be limited by the very low
CO2 uptakes that can be achieved by the tested materials (MSWI-BA and FA),
which implies the requirement of very large amounts of residues in order to achieve
significant uptakes.

Indirect Carbonation

Iizuka et al. [57] proposed to apply pressurised CO2 for carbonating waste cement.
The proposed process consists in a first step in which calcium is extracted from
waste cement in an aqueous solution using pressurised 100 % CO2 and a second
step in which the pressure is reduced causing the extracted calcium to precipitate as
288 R. Baciocchi et al.

calcium carbonate. Experiments showed that up to 50 % of the calcium content


of waste cement could be extracted with a CO2 pressure of 9–30 bar [57].
Kodama et al. [68] developed a process for CO2 sequestration from Ca silicate
industrial residues using the pH swing of a weak base-strong acid solution.
In particular, an ammonium chloride solution was chosen to extract Ca2+ ions
from calcium silicate. As the reaction proceeds, due to NH3 production, the solution
becomes alkaline and, hence, when CO2-containing gas is introduced in the extrac-
tion solution, CaCO3 precipitates and NH4Cl is regenerated. To avoid ammonia
leakage with the flue gas, a NH3 recovery tower was added after the CO2 absorption
unit [68]. Dissolution experiments of various size fractions of crushed converter
steel slag were carried out with a 1 N NH4Cl slurry solution for 1 h at atmospheric
pressure and temperature between 60 and 90  C. A Ca extraction yield of 60 % was
achieved at 80  C for the finest particle size fraction (<63 μm) with a very high Ca
selectivity (>99 %), due to a final solution pH of 9.4, which allowed only Ca2+ to be
dissolved differently from other extracting agents (i.e. HCl and CH3COOH).
Calcium carbonate precipitation conversion showed a maximum conversion of
80 % at 40  C which decreased to 70 % at 80  C, owing to lower CO2 solubility.
A CaCO3 purity of 98 % was obtained, with the formation of calcite at lower
temperature (plate-like morphology) and aragonite at higher temperature (needle-
like morphology) [68].
Another research group investigated the feasibility of enhancing steelmaking
slag dissolution with acetic acid solutions and other chemicals [69, 74–76].
Thermodynamic equilibrium calculations showed that at atmospheric pressure
calcium extraction is exothermic and feasible at temperatures below 156  C,
whereas calcite carbonate precipitation may be obtained at temperatures above
45  C [69]. A high Ca extraction yield was achieved with 33.3 % wt. solutions of
acetic acid for different types of steelmaking slag (BF, AOD, EAF and converter
slag) at 50  C; however, due to the low pH values, other elements dissolved as well,
including Mg, Al, Fe and Si; at 70  C silicon was found to form a gel phase which
could be separated by mechanical filtration and, hence, this temperature was chosen
as the optimum reaction temperature for calcium extraction from steelmaking slag
with acetic acid [69]. Carbonation experiments performed at 30–70  C and 1–30 bar
CO2 on aqueous solutions of calcium extracted from blast furnace slag [74] and
steel converter slag [75] with acetic acid showed that in order for significant calcite
precipitation to occur, the addition of a strong base, such as NaOH, is needed to
increase the pH of the solution. With NaOH addition, calcium carbonate phases
(calcite, aragonite and calcite magnesium) were produced at 30  C and atmospheric
pressure from blast furnace slag [74]. The process scheme developed on the basis of
these results as well as the amounts of reactants required to bind 1 kg CO2 are
reported in Fig. 11.16. Higher purity was achieved with carbonation tests carried
out on aqueous solutions of calcium extracted from converter steel slags dissolved
with weak acetic acid solutions [75]. In this case a maximum calcium conversion to
carbonate of 86 %, with a precipitate purity of up to 99.8 %, was achieved at 30  C
with 10 % vol. CO2. The main limits of this process are the high costs and
requirements of additives such as NaOH and CH3COOH; hence, other solvents
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 289

Fig. 11.16 Process scheme proposed for binding 1 kg of carbon dioxide by carbonating acetate
produced from blast furnace slag (Reprinted from Ref. [74]. Copyright 2008, with permission from
Elsevier)

have been tested for Ca2+ extraction from steelmaking slag [76]. Among these,
ammonium salt solutions were found to be particularly selective in terms of Ca2+
dissolution and operational parameters showed similar effects to those described
by [68].

11.4.3.3 Effects on the Leaching Behaviour of the Treated Residues

As previously mentioned, the effects of accelerated carbonation on the environ-


mental behaviour of industrial residues and in particular on their potential release of
inorganic contaminants such as metals, metalloids and salts has been the focus of
many studies. As reviewed in [67], many works have concerned the effects of
accelerated carbonation as opposed to natural weathering on the leaching behaviour
of MSWI residues by performing different types of tests (e.g. compliance at native
pH, pH dependence or column percolation tests). The multiple reaction mecha-
nisms occurring during carbonation are accompanied by surface interactions
between contaminants and neo-formation minerals and these will impact on the
mobility of trace metals and metalloids in different ways. Hence, the effects exerted
by carbonation on the leaching behaviour of industrial residues are quite complex
and depend on many parameters, such as the specific properties of the individual
trace metals, their solubility curves and behaviour as a function of pH and Eh and
their affinity for minerals as calcite and Al and Fe (hydr)oxides [67]. In general
terms, most studies have reported that carbonation determines a reduction in the pH
of the leachates, which for most types of alkaline residues is above 12, and a
consequent decrease in the mobility of amphoteric elements (such as Cu, Pb and
Zn), whereas various effects on the release of other components such as oxyanion-
forming elements (Cr, Mo and Sb in particular) depend on the characteristics of the
material.
290 R. Baciocchi et al.

Fig. 11.17 Comparison of


compliance leaching test
results for untreated,
carbonated (T ¼ 30  C,
P ¼ 10 bar, L/S ¼ 0.3 l/kg,
t ¼ 2 h) and naturally
weathered RDF-BA
fractions: A ¼ 5.6–12 mm;
B ¼ 2–5.6 mm;
C ¼ 0.425–2 mm;
D ¼ 0.150–0.425 mm;
E < 0.15 mm (Reprinted
from Ref. [42]. Copyright
2010, with permission from
Elsevier)

As an example, Fig. 11.17 reports the effects exerted by accelerated carbonation


on the leaching behaviour of specific size fractions of RDF incineration BA,
resulting from the standardised compliance test (EN 12457-2) [42]. As can be
noted, comparing the results obtained for the accelerated carbonation tests with
those resulting for the same material after 1-year weathering, it may be observed
that generally the accelerated treatment allowed to obtain lower pH values. For the
macro-constituents of the slag, while for Ca leaching the two processes appeared to
exert a very similar effect, Mg showed to be significantly mobilised only for the
carbonated samples. This result may indicate hence that as a result of accelerated
carbonation Mg may have partially formed more soluble phases than the original
Mg silicate mineral constituents of the fresh slag. For amphoteric elements, accel-
erated carbonation appeared to generally exert a stronger effect on reducing the
release of Pb in particular, but also of Cu and Zn. Cr and Mo leaching instead
showed to be enhanced by both types of treatments. It is important to note finally
that no significant variation between the CaCO3 content of untreated and weathered
slag was detected, whereas instead a CO2 uptake of 140 g CO2/kg residues was
obtained for the finest particle size fraction; this suggests, as also indicated in
previous studies (see [67] and references therein), that modifications of the leaching
behaviour of BA do not depend directly on the amount of sequestered CO2.
Similar effects on the leaching behaviour of MSWI-BA were also reported by
Rendek et al. [41], Santos et al. [43], Fernández-Bertos et al. [44], although
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 291

Rendek et al. [41] and Santos et al. [43] reported Cr leaching to decrease upon
carbonation. In this latter study, slurry-phase carbonation was found to be more
effective in reducing the mobility of elements such as Ba and Cu and to cause a
lower mobilisation of V and Sb [43]. Also for APC residues, a significant reduction
in the mobility of Pb, Cu, Zn and Cr was reported [44–46], along with an increase in
Sb release [45, 46]; however, the critical compounds for the management of this
type of residues are soluble salts, chlorides in particular, which do not appear to be
significantly affected by the carbonation treatment.
As for steel slag, carbonation showed to exert a significant decrease in Ca
leaching due to the formation of calcite, which is less soluble compared to (hydr)
oxide and silicate phases, while to increase Mg and Si leaching, consistent with
the pH decrease (for Mg leaching) and with the at least partial dissolution of the
calcium silicate phases [47, 48]. Santos et al. [50] reported a reduction in the
mobility of most trace contaminants (e.g. Mo, Pb, Zn, Ba, Ni, Sb) from AOD and
CC slag, in particular upon slurry-phase carbonation under enhanced conditions.
For V instead, a significant mobilisation effect was reported especially for carbon-
ated BOF samples [47, 49], whereas a slight mobilisation effect for Cr was reported
for EAF and AOD slag [48]. The effects of accelerated carbonation on the leaching
behaviour of pulverised coal fly ash and oil shale ash were first analysed by Reddy
et al. [51, 77]. Treating the humidified samples (20 % moisture) with 100 % CO2 at
around 3 bar for 120 h lowered the pH of the residues from 12.3–12.8 to 8.8–9.7 and
showed to decrease the leaching of Mn, Cd, Pb and Zn for the fly ash and oil shale
samples [51]. In an earlier study, the leaching of F and Mo was indicated to
decrease as a result of accelerated carbonation treatment [77].

11.4.3.4 Effects on the Properties of Products to Use for Specific


Applications

Wet-route carbonation has also been tested as a treatment to manufacture synthetic


aggregates from thermal residues (i.e. MSWI residues, wood ash, paper ash, CBD
and CKD) and waste quarry fines [13]. The process was tested at pilot scale in a unit
designed to produce 100 kg of aggregates/h (see Fig. 11.18). The material is
pelletised in a rotary carbonation reactor at ambient temperature and pressure and
then discharged into a curing chamber circulated with dehumidified CO2 for 7 days.
The produced aggregates showed to present comparable properties to commercially
available lightweight aggregates (e.g. bulk density below 1,000 kg/m3 and a high
water absorption capacity). Concrete blocks were manufactured with carbonated
aggregates and their properties resulted adequate for specific applications, such as
construction of internal partitioning walls [13]. As opposed to the commercially
available fired or sintered aggregates, the proposed process relies on cold bonding
with CO2 and hence should allow to produce carbon-negative aggregates and
construction materials [13].
One of the assets of applying indirect accelerated carbonation routes to minerals
or residues such as steelmaking slag is that these treatments may be applied as a
292 R. Baciocchi et al.

Fig. 11.18 Scheme of the pilot-scale process proposed for aggregates production based on
carbonation of industrial residues (Reprinted from Ref. [13]. Copyright 2009, with permission
from Elsevier)

technique for producing precipitated calcium carbonate (PCC) to use in industrial


applications, e.g. as a filler and coating pigment in plastics, rubbers, paints and
paper manufacturing [69, 74, 75]. PCC is currently produced by carbonating
calcined natural limestone and during calcination more CO2 than that which is
sequestered during the carbonation process is emitted to the atmosphere [69];
hence, the use of calcium silicates for PCC production could contribute in reducing
both CO2 emissions and the consumption of natural resources such as limestone
[75]. Solution temperature is an important parameter for this process since it affects
calcium carbonate morphology [68, 75]; the precipitate formed at low temperature
in fact displays a rhombohedral particle shape and small particle size, which are
considered to be the appropriate characteristics for PCC commercialisation [75].
Recently carbonation of alkaline residues was also tested as a regeneration
treatment for the spent solutions produced by CO2 capture processes based on
absorption with alkali solutions [78, 79]. As discussed in Sect. 11.3.2, the tradi-
tional approach for the regeneration of these solutions is to contact them with
Ca(OH)2 (see Eq. 11.8) that is produced by limestone calcination. Baciocchi
et al. [78] investigated the use of a waste material as calcium hydroxide source so
to reduce the consumption of alkali reagents and improve the economical and
environmental performance of the alkali absorption process proposed as a biogas-
upgrading technique with the additional benefit of permanently storing the sepa-
rated CO2 as calcite. On the basis of the results of preliminary lab-scale tests [78],
a pilot-scale regeneration unit employing APC residues was designed, built and
installed in a landfill site downstream of an absorption column treating landfill
biogas with KOH/NaOH solutions [79]. Results of the combined absorption and
regeneration pilot tests showed overall regeneration efficiencies of the alkali solu-
tions of around 50–60 %. Furthermore, the performance of the absorption step
appeared not to be affected by the use of regenerated solutions, while the solid
product of the process, presenting a calcite content of over 80 %, presented an
improved leaching behaviour compared to the raw APC residues [78, 79]. However,
one of the critical aspects for the applicability of this process is related to the limited
availability of Ca(OH)2-rich residues such as APC residues.
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 293

11.4.4 Energy and Material Requirements of Accelerated


Carbonation

Although the feasibility of accelerated carbonation of minerals and industrial residues


was proven at lab scale, the significant requirement of materials and energy associated
to the process has hindered so far its application at larger scale. Recently, several
studies have focused on the evaluation of the energy penalties related to mineral
carbonation adopting the slurry-phase route scheme [29, 80–82]. However, the energy
penalties evaluated in these works differ considerably from one study to the other,
even considering the different power plant types used as reference and also when the
evaluation was based on the same experimental data, probably as a consequence of the
differences in the assumptions made and in the selected process layout.
As discussed in Sect. 11.4.2, the Albany Research Center (ARC) presented a
preliminary study aimed at developing an ex situ aqueous process to convert
magnesium silicate-rich ultramafic rocks and minerals, such as olivine, wollastonite
and serpentine, to magnesium carbonates by using gaseous CO2 in an aqueous
solution [80]. In this study seven kinds of ultramafic minerals were analysed
considering various pretreatment options and operating conditions for the carbon-
ation reaction and based on the results of lab-scale tests, the feasibility of the
process was investigated. At the optimal operating conditions using olivine (L/S
¼ 2.33 kg/kg, t ¼ 120 min, T ¼ 185  C, pCO2 ¼ 150 bar and d < 37 μm), the
authors determined an energy requirement of 352 MW for the storage of 100 % of
the CO2 emissions from a 1.3 GW coal-fired electric generation plant characterised
by a heat release of around 10 GJ/tCO2 emitted. The energy demand of this process
using olivine, expressed in terms of CO2 emissions, was estimated to be around
28 % of the total amount of carbon dioxide sequestered that corresponds to an
energy requirement of 2.66 GJ/t CO2 [80].
Huijgen et al. [29] evaluated the energetic performance of CO2 sequestration by
aqueous carbonation of Ca silicate ores and industrial residues. Specifically they
considered data obtained from carbonation experiments carried out at lab scale
using both a wollastonite and steel slag feedstock and on this basis they assumed a
mineral carbonation process layout, thus determining the CO2 sequestration effi-
ciency at various operating conditions. The minimum energy penalties in terms of
CO2 emissions determined for wollastonite and steel slag were 16 % and 17 %,
respectively (at L/S ¼ 5 kg/kg, t ¼ 15 min, T ¼ 473 K, pCO2 ¼ 2 MPa, particle
size lower than 38 μm), making reference to a power plant characterised by a heat
release of 18 GJ/tCO2 emitted. For an L/S ¼ 2 kg/kg, higher energy penalties
resulted for wollastonite and in particular for steel slag (i.e. 25 % and 31 %,
respectively). The grinding of the feedstock and the compression of the carbon
dioxide were identified as the main energy-consuming process steps. In addition, a
sensitivity analysis has shown that further grinding (particularly for wollastonite) or
reducing the CO2 partial pressure (in the case of steel slag) can potentially improve
the process efficiency [29].
Kelly et al. [81] performed a preliminary energy balance for three CO2
mineralisation pathways at industrial scale, i.e. industrial caustics, naturally
294 R. Baciocchi et al.

occurring minerals and industrial wastes. The basis for this evaluation is a
theoretical 1 GW coal-fired power plant emitting 8  106 t CO2/y (i.e. a heat
release of around 10 GJ/tCO2 emitted), which is captured by one of the considered
mineral carbonation pathways. The estimation of the energy penalties for the
carbonation of olivine and wollastonite was performed on the basis of previous
works [29, 80] taking also into account the requirements for mining and CO2
separation. The resulting energy penalties ranged between 55 and 69 % for olivine
and were over 100 % for wollastonite [81]. Regarding industrial residues, Kelly
et al. [81] concluded that the energy penalty associated to CO2 mineralisation of
steel slag proposed by Huijgen et al. [29] is greater than 100 %, making this process
unlikely to be feasible. On the other hand, the CO2 mineralisation process with fly
ash proposed by Reddy et al. [73] was considered feasible, as it presented a much
lower energy penalty (9–22 %). It is worth pointing out that the energy penalties
calculated by Kelly et al. [81] for steel slags were estimated assuming no energy
recovery; however, as suggested also by Huijgen et al. [29], it is reasonable to
assume that the heat of the water-mineral slurry can be recovered allowing to
greatly reduce the energy penalties estimated by Kelly et al. [81].
Recently Kirchofer et al. [82] estimated the energy and material requirements of
different aqueous carbonation processes using both natural silicate minerals
(i.e. olivine and serpentine) and industrial by-products, such as cement kiln dust
[59], coal fly ash [54] and steelmaking slag [29], in order to compare their
environmental performance by life cycle assessment (LCA). The minimum energy
penalty expressed in terms of percent CO2 emitted per CO2 sequestered, consider-
ing natural gas as energy source, was obtained for cement kiln dust at ambient
temperature and pressure conditions, resulting equal to 14 %, while the results
obtained for the other considered materials were 25 % for olivine, 75 % for
serpentine, 34 % for steel slag and 45 % for fly ash. Furthermore, within the
range of the considered operating conditions, heating provided the most relevant
contribution, followed by mixing and grinding. It should be noted that in this study
the energy requirements for CO2 capture were not considered while those related to
mining and transport of reactants and products were accounted for. Regarding the
carbonation of minerals, in this study the energy penalties were estimated on the
basis of the experimental results obtained in dissolution tests, assuming that all
the dissolved cations react with CO2 in the precipitation stage. This may explain the
observed differences in the effects of the operating conditions and resulting energy
requirements compared to the previously mentioned studies. As for the carbonation
of residues, it is worth mentioning that the same process scheme was applied for all
of the tested materials regardless of the applied L/S ratio.

11.5 Conclusions and Perspectives

Accelerated carbonation has been investigated for more than a decade as an option
for carbon capture and storage. Apart from the well-established hot carbonate
process, that has found an interesting market application in the ammonia synthesis
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 295

sector, the calcium looping technology seems to be the most promising and
developed application of accelerated carbonation for carbon capture. Nevertheless,
the process still suffers from the low stability of the calcium-based sorbents, whose
CO2 carrying capacity decreases sharply after a few carbonation/calcination cycles.
Besides, the effective advantage of this process in terms of energy requirements
with respect to commercial ethanolamine processes must still be demonstrated.
Differently, the perspectives for the alkali absorption process are quite negative, in
view of the very large energy inputs required to regenerate the solvent. The only
solution to this issue may come from the use of alternative alkalinity sources, such
as industrial residues rich in calcium hydroxide, that may be used for the solvent
regeneration phase in place of raw calcium hydroxide; anyhow, the availability of
these residues represents the main bottleneck for the deployment of this technology.
Accelerated carbonation as storage option has also been the subject of intense
research efforts aimed at improving the efficiency of the process. So far, the direct
aqueous route still represents the benchmark technology for mineral carbonation.
Unfortunately, this route suffers from too high energy requirements, mostly related
to the mechanical and thermal activation of the minerals and to the heat required to
bring the mineral slurry to the optimal reaction temperature. The indirect route in
principle could allow to reduce the energy requirements of the process; neverthe-
less, its feasibility is questionable due to the consumption of chemicals needed in
these multistep processes and given the uncertainty on the effective possibility of
recycling them. Thus summarising, the energy penalties associated to mineral
carbonation would be of the same order of magnitude of the ones associated to
carbon capture. Therefore, the only feasible perspective for the implementation of
this technology seems to be represented by the application of mineral carbonation
directly to the emission source, such as flue gas or syngas. Lumping capture and
storage in a single step could probably reduce the overall energy penalty, allowing
to fix CO2 directly in a stable and definitive manner.
Differently from mineral carbonation, the many kinds of alkaline residues
produced from thermal processes, steelmaking and cement industries allowed for
the development of diverse strategies and reaction routes. Based on the overview
made in this chapter, carbonation of industrial residues may represent a suitable
option for storing at least part of the carbon dioxide emissions produced by these
industrial sectors. These materials are typically more reactive than silicate minerals,
allowing to achieve reasonable extents of carbonation at milder operating condi-
tions. Besides, as these materials are already available at the CO2 point source, all
issues related to mining and transport would be avoided. Anyhow, also in this case
the energy requirements associated to carbonation are not negligible. The wet route,
at mild operating temperature and pressure, using a fairly concentrated CO2 source,
such as syngas or biogas, seems the most promising option for the application of the
process.
Finally, the role of accelerated carbonation in CCUS must be briefly discussed.
The term carbon utilisation suggests that some of the anthropogenically emitted
CO2 can be used for different purposes, but without any guarantee on the final
destiny of the used carbon. On the other hand, the term storage means that the
296 R. Baciocchi et al.

anthropogenic CO2 is isolated or fixed definitively. Given these definitions, as


accelerated carbonation will always lead to the permanent fixation of some CO2,
we observe that it should be included among the storage options. The term
utilisation could be eventually coupled to the term storage only when the storage
capacity is negligible with respect to the emission source that we want to address
and the main benefit of the process is given by the improvement of the environ-
mental properties of the material or by the manufacturing of a product that can be
employed for different applications.

References

1. Boschi C, Dini A, Dallai L et al (2010) Enhanced CO2-mineral sequestration by cyclic


hydraulic fracturing and Si-rich fluid infiltration into serpentinites at Malentrata (Tuscany,
Italy). Chem Geol 265:209–226
2. Lackner KS, Wendt CH, Butt D et al (1995) Carbon dioxide disposal in carbonate minerals.
Energy 20:1153–1170
3. Haug TA, Kleiv RA, Munz IA (2010) Investigating dissolution of mechanically activated
olivine for carbonation purposes. Appl Geochem 25:1547–1563
4. Seifritz W (1990) CO2 disposal by means of silicates. Nature 345:486
5. Abanades JC, Alvarez D (2003) Conversion limits in the reaction of CO2 with lime. Energy
Fuels 17:308–315
6. Yi F, Zou HK, Chu GW et al (2009) Modeling and experimental studies on absorption of CO2
by Benfield solution in rotating packed bed. Chem Eng J 145:377–384
7. IPCC (2005) IPCC special report on carbon dioxide capture and storage. In: Metz B,
Davidson O, de Coninck HC, Loos M, Meyer LA (eds) Prepared by Working Group III of
the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge/
New York, 442 pp
8. Lackner K, Duby P, Yegulalp T et al (2008) AISI/DOE technology roadmap – TRP9957
Integrating steel production with mineral carbon sequestration. US DOE Technical Report,
New York, doi: 10.2172/938528
9. Lackner KS (2002) Carbonate chemistry for sequestering fossil carbon. Annu Rev Energy
Environ 27:193–232
10. Zevenhoven R, Teir S, Eloneva S (2006) Chemical fixation of CO2 in carbonates: routes to
valuable products and long-term storage. Catal Today 115:73–79
11. Gunning PJ, Hills CD, Carey PJ (2010) Accelerated carbonation treatment of industrial
residues. Waste Manag 30:1081–1090
12. Kirchofer A, Becker A, Brandt A et al (2013) CO2 mitigation potential of mineral carbonation
with industrial alkalinity sources in the United States. Environ Sci Technol 47:7548–7554
13. Gunning PJ, Hills CD, Carey PJ (2009) Production of lightweight aggregate from industrial
waste and carbon dioxide. Waste Manag 29:2722–2728
14. Shimizu T, Hirama T, Hosoda H et al (1999) A twin fluid-bed reactor for removal of CO2 from
combustion processes. Trans IChemE 77:62–68
15. Blamey J, Anthony EJ, Wang J et al (2010) The calcium looping cycle for large-scale CO2
capture. Prog Energy Combust Sci 36:260–279
16. Zeman F (2008) Effect of steam hydration on performance of lime sorbent for CO2 capture. Int
J Greenh Gas Control 2:203–209
17. Abanades JC, Rubin ES, Anthony EJ (2004) Sorbent cost and performance in CO2 capture
systems. Ind Eng Chem Res 43:3462–3466
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 297

18. Gupta H, Fan LS (2002) Carbonation-calcination cycle using high reactivity calcium oxide for
carbon dioxide separation from flue gas. Ind Eng Chem Res 41:4035–4042
19. Stendardo S, Andersen LK, Herce C (2013) Self-activation and effect of regeneration
conditions in CO2–carbonate looping with CaO–Ca12Al14O33 sorbent. Chem Eng J
220:383–394
20. Sanyal D, Vasishtha N, Saraf DN (1988) Modeling of carbon dioxide absorber using hot
carbonate process. Ind Eng Chem Res 27:2149–2156
21. Nagasawa H, Yamasaki H, Iizuka A et al (2009) A new recovery process of carbon dioxide
from alkaline carbonate solution via electrodialysis. AIChE J 55:3286–3293
22. Baciocchi R, Storti G, Mazzotti M (2006) Process design and energy requirements for the
capture of carbon dioxide from air. Chem Eng Process 34:1047–1058
23. APS (2011) Direct air capture of CO2 with chemicals: a technology assessment for the APS
panel on public affairs. American Physical Society Technical Report, Washington DC, April
24. Aroonwilas A, Tontiwachwuthikul P (2000) Mechanistic model for prediction of structured
packing mass transfer performance in CO2 absorption with chemical reactions. Chem Eng Sci
55:3651–3663
25. Zevenhoven R, Teir S, Eloneva S (2008) Heat optimisation of a staged gas–solid mineral
carbonation process for long-term CO2 storage. Energy 33:362–370
26. Nduagu E, Biörklöf T, Fagerlund J et al (2012) Production of magnesium hydroxide from
magnesium silicate for the purpose of CO2 mineralization – Part 2: Mg extraction modeling
and application to different Mg silicate rocks. Miner Eng 30:87–94
27. Gerdemann SJ, O’Connor WK, Dahlin DC et al (2007) Ex situ aqueous mineral carbonation.
Environ Sci Technol 41:2587–2593
28. Huijgen WWJ, Witkamp GJ, Comans RNJ (2006) Mechanisms of aqueous wollastonite
carbonation as a possible CO2 sequestration process. Chem Eng Sci 61:4242–4251
29. Huijgen WWJ, Ruijg GJ, Comans RNJ et al (2006) Energy consumption and net CO2
sequestration of aqueous mineral carbonation. Ind Eng Chem Res 45:9184–9194
30. Huijgen WWJ, Comans RNJ, Witkamp G (2007) Cost evaluation of CO2 sequestration by
aqueous mineral carbonation. Energy Conv Manag 48:1923–1935
31. Hänchen M, Prigiobbe V, Storti G et al (2006) Dissolution kinetics of fosteritic olivine at
90–150  C including effects of the presence of CO2. Geochim Cosmochim Acta 70:4403–4416
32. Prigiobbe V, Costa G, Baciocchi R (2009) The effect of CO2 and salinity on olivine dissolution
kinetics at 120  C. Chem Eng Sci 64:3510–3515
33. Park AAH, Fan LS (2004) CO2 mineral sequestration: physically activated dissolution of
serpentine and pH swing process. Chem Eng Sci 59:5241–5247
34. Teir S, Kuusik R, Fogelholm CJ et al (2007) Production of magnesium carbonates from
serpentinite for long-term storage of CO2. Int J Miner Process 85:1–15
35. Wang X, Maroto-Valer MM (2011) Dissolution of serpentine using recyclable ammonium
salts for CO2 mineral carbonation. Fuel 90:1229–1237
36. Hänchen M, Prigiobbe V, Baciocchi R et al (2008) Precipitation in the Mg-carbonate system—
effects of temperature and CO2 pressure. Chem Eng Sci 63:1012–1028
37. Bobicki ER, Liu Q, Xu Z et al (2012) Carbon capture and storage using alkaline industrial
wastes. Prog Energy Combust 38:302–320
38. Pan SY, Chang EE, Chiang PC (2012) CO2 capture by accelerated carbonation of alkaline
wastes: a review on its principles and applications. Aerosol Air Qual Res 12:770–791
39. Huijgen WJJ, Comans RNJ (2003) Carbon dioxide sequestration by mineral carbonation:
literature review. Technical Report ECN-C—03016, Petten, February
40. Van Gerven T, Van Keer E, Arickx S et al (2005) Carbonation of MSWI-bottom ash to
decrease heavy metal leaching, in view of recycling. Waste Manag 25:291–300
41. Rendek E, Ducom G, Germain P (2006) Carbon dioxide sequestration in municipal solid waste
incinerator (MSWI) bottom ash. J Hazard Mater B128:73–79
42. Baciocchi R, Costa G, Lategano E et al (2010) Accelerated carbonation of different size
fractions of bottom ash from RDF incineration. Waste Manage 30:1310–1317
298 R. Baciocchi et al.

43. Santos RM, Mertens G, Salman M et al (2013) Comparative study of ageing, heat treatment
and accelerated carbonation for stabilization of municipal solid waste incineration bottom ash
in view of reducing regulated heavy metal/metalloid leaching. J Environ Manage 128:807–821
44. Fernández-Bertos M, Li X, Simons SJR et al (2004) Investigation of accelerated carbonation
for the stabilization of MSW incinerator ashes and the sequestration of CO2. Green Chem
6:428–436
45. Baciocchi R, Costa G, Di Bartolomeo E et al (2009) The effects of accelerated carbonation on
CO2 uptake and metal release from incineration APC residues. Waste Manage 29:2994–3003
46. Cappai G, Cara S, Muntoni A (2012) Application of accelerated carbonation on MSW
combustion APC residues for metal immobilization and CO2 sequestration. J Hazard Mater
207–208:159–164
47. Huijgen WJJ, Comans RNJ (2006) Carbonation of steel slag for CO2 sequestration: leaching of
products and reaction mechanisms. Environ Sci Technol 40:2790–2796
48. Baciocchi R, Costa G, Bartolomeo D et al (2010) Carbonation of stainless steel slag as a
process for CO2 storage and slag valorization. Waste Biomass Valor 1:467–477
49. van Zomeren A, van der Laan SR, Kobesen HBA et al (2011) Changes in mineralogical and
leaching properties of converter steel slag resulting from accelerated carbonation at low CO2
pressure. Waste Manage 31:2236–2244
50. Santos RM, Van Bouwel J, Vandevelde E et al (2013) Accelerated mineral carbonation of
stainless steel slags for CO2storage and waste valorization: effect of process parameters on
geochemical properties. Int J Greenh Gas Control 17:32–45
51. Reddy KJ, Drever JI, Hausfurther VR (1991) Effects of a CO2 pressure process on the
solubilities of major and trace elements in oil shale solid wastes. Environ Sci Technol
25:1466–1469
52. Uibu M, Uus M, Kuusik R (2009) CO2 mineral sequestration in oil-shale wastes from Estonian
power production. J Environ Manage 90:1253–1260
53. Arickx S, Van Gerven T, Vandecasteele C (2006) Accelerated carbonation for treatment of
MSWI bottom ash. J Hazard Mater B137:235–243
54. Back M, Kuehn M, Stanjek H et al (2008) Reactivity of alkaline lignite fly ashes towards CO2
in water. Environ Sci Technol 42:4520–4526
55. Bonenfant D, Kharoune L, Sauvé S et al (2008) CO2 sequestration by aqueous red mud
carbonation at ambient pressure and temperature. Ind Eng Chem Res 47:7617–7622
56. Fernández-Bertos M, Simons SJR, Hills CD et al (2004) A review of accelerated carbonation
technology in the treatment of cement-based materials and sequestration of CO2. J Hazard
Mater B112:193–205
57. Iizuka A, Fujii M, Yamasaki A et al (2004) Development of a new CO2 sequestration process
utilizing the carbonation of waste cement. Ind Eng Chem Res 43:7880–7887
58. Huijgen WJJ, Witkamp GJ, Comans RNJ (2005) Mineral CO2 sequestration by steel slag
carbonation. Environ Sci Technol 39:9676–2682
59. Huntzinger DN, Gierke JS, Kawatra SK et al (2009) Carbon dioxide sequestration in Cement
Kiln Dust through mineral carbonation. Environ Sci Technol 43:1986–92
60. Jia L, Anthony EJ (2000) Pacification of FBC ash in a pressurized TGA. Fuel 79:1109–1114
61. Baciocchi R, Polettini A, Pomi R et al (2006) CO2 sequestration by direct gas-solid carbon-
ation of APC residues. Energy Fuels 20:1933–1940
62. Prigiobbe V, Polettini A, Baciocchi R (2009) Gas–solid carbonation kinetics of air pollution
control residues for CO2 storage. Chem Eng J 148:270–278
63. Bonenfant D, Kharoune L, Sauvé S et al (2008) CO2 sequestration potential of steel slags at
ambient pressure and temperature. Ind Eng Chem Res 47:7610–7616
64. Montes-Hernandez G, Pérez-López R, Renard F et al (2009) Mineral sequestration of CO2 by
aqueous carbonation of coal combustion fly-ash. J Hazard Mater 161:1346–1354
65. Young JF, Berger RL, Breese J (1974) Accelerated curing of compacted calcium silicate
mortars on exposure to CO2. J Am Ceram Soc 57:394–397
66. Papadakis VG, Vayenas CG, Fardis MN (1991) Experimental investigation and mathematical
modelling of the concrete carbonation problem. Chem Eng Sci 46:1333–1338
11 Accelerated Carbonation Processes for Carbon Dioxide Capture. . . 299

67. Costa G, Baciocchi R, Polettini A et al (2007) Current status and perspectives of accelerated
carbonation processes on municipal waste combustion residues. Environ Monit Assess
135:55–75
68. Kodama S, Nishimoto T, Yamamoto N et al (2008) Development of a new pH-swing CO2
mineralization process with a recyclable reaction solution. Energy 33:776–784
69. Teir S, Eloneva S, Fogelholm CJ et al (2007) Dissolution of steelmaking slags in acetic acid for
precipitated calcium carbonate production. Energy 32:528–539
70. Johnson DC, Macleod CL, Carey PJ et al (2003) Solidification of stainless steel slag by
accelerated carbonation. Environ Technol 24:671–678
71. Pérez-López R, Montes-Hernandez G, Nieto JM et al (2008) Carbonation of alkaline paper
mill waste to reduce CO2 greenhouse gas emissions into the atmosphere. Appl Geochem
23:2292–2300
72. Mostbauer P, Lenz S, Lechner P (2008) MSWI bottom ash for upgrading of biogas and landfill
gas. Environ Technol 29:757–764
73. Reddy KJ, Kohn S, Weber H et al (2011) Simultaneous capture and mineralization of coal
combustion flue gas carbon dioxide (CO2). Energy Proced 4:1574–83
74. Eloneva S, Teir S, Salminen J et al (2008) Fixation of CO2 by carbonating calcium derived
from blast furnace slag. Energy 33:1461–1467
75. Eloneva S, Teir S, Salminen J et al (2008) Steel converter slag as raw material for precipitation
of pure calcium carbonate. Ind Eng Chem Res 47:7104–7111
76. Eloneva S, Mannisto P, Said A et al (2011) Ammonium salt-based steelmaking slag carbon-
ation: precipitation of CaCO3 and ammonia losses assessment. Greenh Gas Sci Technol
1:305–311
77. Reddy KJ, Gloss SP, Wang L (1994) Reaction of CO2 with alkaline solid wastes to contam-
inant mobility. Water Res 28:1377–1382
78. Baciocchi R, Costa G, Gavasci R et al (2012) Regeneration of a spent alkaline solution from a
biogas upgrading unit by carbonation of APC residues. Chem Eng J 179:63–71
79. Baciocchi R, Carnevale E, Corti A et al (2013) Innovative process for biogas upgrading with
CO2 storage: results from pilot plant operation. Biomass Bioenergy 53:128–137
80. O’Connor WK, Dahlin DC, Rush GE et al (2005) Aqueous mineral carbonation: mineral
availability, pretreatment, reaction parametrics and process studies. Report DOE/ARC-TR-04-
002. Albany Research Center, Albany
81. Kelly KE, Silcox GD, Sarofim AF et al (2011) An evaluation of ex situ, industrial-scale,
aqueous CO2 mineralization. Int J Greenh Gas Control 5:1587–1595
82. Kirchofer A, Brandt A, Krevor S et al (2012) Impact of alkalinity sources on the life-cycle
energy efficiency of mineral carbonation technologies. Energy Environ Sci 5:8631–8641
Part III
Biological Reactions
Chapter 12
Carbon Dioxide Sequestration
by Biological Processes

Kanhaiya Kumar and Debabrata Das

12.1 Introduction

Carbon dioxide is one among many greenhouse gases such as methane, ozone, NOx,
and water vapor. Greenhouse gases act as blanket, which retain the outgoing sun’s
heat (infrared rays) into the earth’s atmosphere. In absence of greenhouse gases,
earth would be colder. Contrary to this, increase in any one of the greenhouse gases
will result in extra trapping of heat, causing global warming. CO2 constitutes very
small portion of the gases present in the earth’s atmosphere. Natural range of CO2
concentration over the last 650,000 years was in the range of 180–300 ppmv.
However, presently CO2 concentration has crossed the natural range and reached
as high as 397.34 ppmv in March 2013. Moreover, the Keeling curve reveals
progressively faster rise in CO2 concentration after industrialization [1]. Rate of
increase in atmospheric CO2 was 1.94 ppm/year in 2011, which was more than
twice the estimated value in 1959 [1]. Temperature contributes significantly to
make planet habitable for the living beings. Therefore, trace amount of CO2 in the
earth’s atmosphere is required for maintaining stable temperature. Increase or
decrease of few degrees of global temperature can have a devastating effect on
earth. Global atmosphere and ocean temperature have increased by 0.6  C and
0.3  C, respectively, in the last century despite the fact that solar output witnessed
decline in the same period [2]. Rise in temperature is the reason for climate change
and melting of glaciers, thus causing rise in the sea level. In addition, global
warming will cause increase in soil microbe’s respiration and further addition of
greenhouse gas CO2 [3]. In the last century, global sea level rose by 17 cm
[2]. Nearly 2 billion tons of CO2 is being absorbed in the upper layer of ocean
per year. Increase in CO2 absorption resulted in the increase of acidification that
adversely affects the aquatic life [2]. The International Panel on Climate Change

K. Kumar • D. Das (*)


Department of Biotechnology, Indian Institute of Technology Kharagpur,
I.I.T. Kharagpur, India
e-mail: ddas.iitkgp@gmail.com

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 303
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_12,
© Springer-Verlag Berlin Heidelberg 2014
304 K. Kumar and D. Das

(IPCC) forecast increase in CO2 concentration up to 570 ppmv by the end of the
twenty-first century. This will cause nearly a rise of 1.9  C of mean global
temperature, causing an increase in mean sea level by 38 cm [4]. It is disputable
to link rise in temperature with increase in CO2 concentration. Some reports suggest
present temperature rise is the result of natural cycle of rise and fall in temperature
over the earth’s history. However, some researchers believe present rise in temperature
is due to an increase in CO2 concentration. It becomes more important as some reports
reveal unprecedented rate of increase in temperature is contrary to natural cycle of
earth temperature observed so far [2].
CO2 is being injected into the atmosphere from natural as well as artificial
sources. Natural sources of CO2 are volcanic eruptions, decomposition of organic
matters, and autotrophic and heterotrophic respiration [5]. However, there is a
natural mechanism of CO2 removal such as oceanic and terrestrial CO2 removal
from the atmosphere. Imbalance in natural CO2 addition and removal from the
atmosphere is due to anthropogenic emission of CO2 by human activity despite the
fact that natural processes remove 50 % of the anthropogenic CO2 emissions.
Therefore, increasing consumption of fossil fuels is the main matter of concern
due to the anthropogenic CO2 emissions. Total annual anthropogenic CO2 emission
due to fossil fuel consumption is 29 Gt per year [6]. Today, power plants, cement,
steel manufacturing industries, transportation, and household usages are dependent
on fossil fuels [7]. Coal-fired power plants consume nearly one third of the total
fossil fuel consumption and the remaining from fossil fuel usages in other sectors
such as transportation, industry, and homes [2, 8]. Increase in human population
and modernization are the reasons for the booming of these sectors.
In this scenario, it is necessary to adopt some strategies to sequester CO2 from
the earth’s atmosphere. CO2 sequestration strategies can be divided into abiotic and
biotic categories. Abiotic categories involve scrubbing, mineral carbonation, and
geological and ocean injection. Several techniques are available to separate CO2
from the flue gas. Further, they can be stored at various locations such as oceans,
deep aquifers, and depleted oil and gas reservoirs [9]. However, these methods can
pose potential threat for safety and environmental impact due to accidental leakage
especially in long term [10]. Biotic categories include CO2 sequestration in oceanic,
terrestrial, and secondary carbonates. There are several natural phenomena occur-
ring in the ocean, which ultimately remove atmospheric CO2 [11]. Oceanic CO2 at
the surface is in equilibrium with atmospheric CO2. CO2 dissolved in water forms
weak acid, which reacts with carbonate anions and water to form bicarbonate.
Continuous supplies of cations are required to maintain the buffering capacity of
bicarbonate system. CO2 solubility gradient is another phenomenon by which large
amount of CO2 is being sent to the bottom of ocean. CO2 is more soluble in cold and
saline water. Therefore, cold dense water masses at higher latitude especially at the
North Atlantic and Southern Ocean confluence sink into the interior of ocean
carrying huge amount of CO2 [11]. This eventually gets trapped by less dense
water at the top for several hundred years. Further, oceanic phytoplanktons play a
vital role by absorbing atmospheric CO2 at the surface of the ocean during photo-
synthesis. Nearly 25 % of carbon fixed by these processes sinks to the bottom of the
12 Carbon Dioxide Sequestration by Biological Processes 305

ocean. It has been estimated that 11–16 Gt of carbon is removed from the atmo-
sphere by this process [11]. Oceanic phytoplanktonic photosynthesis lowers the
atmospheric CO2 by 150–200 ppmv [11]. CO2 is also absorbed naturally such as
through sedimentation. CO2 conversion into bicarbonate rocks is a natural process.
Nearly 2 Gt of carbon is sequestered in the soil through burial of decomposed
organic material from plants, animals, biomass, and agriculture by microbes
[6]. However, most of those are released back into the atmosphere during soil
erosion and oxidation. Biochar is another promising material helpful in CO2
removal. Biochar is charcoal, which is produced when smoldering biomass is
burnt in limited oxygen producing little heat. The advantage of biochar in CO2
sequestration is that it gives stability to carbon present in the biomass against
microbe’s oxidation [6] and, thus, prevents the release of CO2 into the atmosphere.
Land-based plants consume CO2 in photosynthesis and release CO2 during
respiration. However, there is net CO2 sequestration from the atmosphere. Plants
are not capable of utilizing the concentrated CO2 present in flue gas. It has been
estimated that afforesting 22 % of earth’s terrestrial surface, which is equal to
present forested land, is required to sequester CO2 emitted due to fossil fuel usages
[12]. Other reports say only 345 million hectares is available for afforestation, which
will result in 1.5 Gt C annually [6]. This can lead to little less than 20 % reduction in
anthropogenic CO2 emission. However, afforestation in total available area is not
possible due to huge pressure on land for other purposes. Each year nearly 10 % of
total atmospheric CO2 is fixed into carbohydrates by the photosynthetic process
[13]. Total global CO2 removal by land-based plants and oceanic phytoplanktons
are 403 Gt CO2 (equivalent to 110 Gt carbon) and 385 Gt CO2, respectively
[8]. Microalgae, seaweeds, and higher aquatic plants are other alternatives for CO2
sequestration because of their photosynthetic ability. However, the latter two are not
promising because of low CO2 and other nutrient mass transfer [14]. They require
high amount of energy input for mixing. However, seaweeds are more promising
compared to higher aquatic plants because it is difficult to achieve proper water
exchange at the submerged plants and their dense stands [14]. Therefore, seaweeds
are grown near the seashore and shallow ocean systems. Carbon is provided by
seawater exchange system but requires high amount of energy for water pumping.
Use of microalgae, cyanobacteria, and other bacteria is another alternative
available for CO2 sequestration. A large number of microbes are known capable
of utilizing CO2. Except microalgae, all require some inorganic reducing agent such
as H2, H2S, NH3, etc. [14]. Photosynthetic efficiency of algae is 10 times higher
than that of terrestrial plants. They are efficient in growing at relative higher
concentration of CO2 compared to plants. This makes them a suitable candidate
for CO2 sequestration from the flue gas. Cultivation of algae requires less water, and
they can be grown in non-fertile land. Further, their biomass can be used for food
and feed supplements, for the production of biofuels such as biohydrogen,
bioethanol, biodiesel, and industrially important biomolecules and biofertilizers
[7, 15, 16].
Use of carbonic anhydrase (CA) is another thrust area for CO2 sequestration.
Carbon content of earth’s lithosphere is 42 % w/w in the form of CaCO3 and other
306 K. Kumar and D. Das

carbonates which indicates transformation of gaseous CO2 into solid carbonates is a


geological stable process and has potential to exploit for CO2 sequestration
[17]. One of the reactions, hydration of CO2 to carbonic acids, is the rate limiting
step in CO2 mineralization reaction [18]. CA is the catalyst of this reaction, and
the use of this enzyme has been found to enhance the rate of reaction mani-
fold [19]. Advantages of using CA for CO2 sequestration are eco-friendliness,
cost-effectiveness, simple process, and abundance of CA enzymes among
microorganisms [20].
Thus, the present study aims to summarize various biological processes such as
use of plants, microalgae, cyanobacteria, and other non-photosynthetic microbes
for CO2 sequestration. It also attempts to highlight different pathways occurring in
biological systems along with the use of CA for this cause.

12.2 Biological Processes of CO2 Sequestration

12.2.1 CO2 Sequestration by Plants

The amount of CO2 required for the terrestrial plants and crops is very less, nearly
equal to atmospheric CO2 concentration. Limited number of reports are available,
demonstrating the long-term effect of CO2 concentration on plants because of high
amount of time and space required along with variation in experimental conditions.
Therefore, it is not clear that the rise in CO2 concentration will increase the CO2
sequestration in existing forests [21]. Most suitable CO2 concentration for plants
obtained in greenhouses was nearly three times (0.1 %) the atmospheric CO2. An
experiment conducted by Norby et al. (2005) by analyzing four Free Air Carbon
dioxide Enrichment (FACE) studies on forest reported that there will be 23 %
increase in the plant productivity at the predicted level of CO2 in 2050 compared to
that of the present atmospheric CO2 concentration [22]. However, they cautioned
that increment in productivity may not increase the long-term CO2 sequestration.
Most of the trees in forest (~95 % of all higher plants) are C3 type having positive
photosynthetic response to elevated CO2 concentration [21, 23, 24]. Nearly one
third of the global land is covered by forest. In another report, on average, 60 % and
27 % enhancement of photosynthesis and growth, respectively, for trees have been
reported [25]. Elevated CO2 concentration decreases the stomata aperture opening
and reduces the water loss. Trees planted at CO2 spring in Italy were found growing
at similar rate compared to that of normal atmospheric CO2 concentration [26]. Res-
piration rate has been found to decline in elevated CO2 environment with little
improvement in the photosynthetic ability of the plants. Thus, it is impractical to
use terrestrial plants and crops for CO2 sequestration from highly rich CO2 stream
of flue gas emitted from power plants [14]. The effect of temperature, one of the
associated parameters with CO2 concentration, has also been studied. Net CO2
removal rate by plants was found to increase with temperature [27]. Further
increase in temperature from optimum resulted in decreased CO2 sequestration
12 Carbon Dioxide Sequestration by Biological Processes 307

rate. Therefore, CO2 sequestration rate by green plants varies from day to day and
season to season. It was observed that in CO2-enriched atmosphere, plants shift the
optimum temperature of CO2 sequestration to higher level. Moreover, at this
condition, process was found to be less sensitive to increase in temperature
[27]. Despite the fact that plants have limited applicability in CO2 sequestration
at highly CO2-rich environment, afforestation has a number of advantages. Firstly,
nearly 30–40 % of captured carbon from atmosphere gets stored in depth of soil
system through plant roots [28]. Secondly, cultivated agricultural crops contribute
large amount of atmospheric CO2 removal (in the order of 190 t ha1 C) into the soil
system acting as perpetual sink [28]. Thirdly, the lack of proper nitrogen fertilizer
for land-based plants is one of the limiting factors. Therefore, CO2 removal rate by
plants can be enhanced significantly by the use of nitrogen fertilizer [8].

12.2.1.1 Mechanism of CO2 Fixation

Each leaf cell contains nearly 50 chloroplasts [13]. Stomata are the entry sites of
CO2 into the leaf where it interacts with the RuBisCO present in chloroplasts and
gaseous CO2 reduced into carbohydrates. Stomata of leaves have been found
unresponsive to higher CO2 concentration. RuBisCO is the key enzyme of the
Calvin cycle for all the three types of the plant C3, C4, and CAM. Km of CO2 for the
C3 plants is 15–25 μM, and carboxylase activity of RuBisCO operates below the
Km of CO2 and is not more than 25–30 % (equilibrium concentration of CO2 in
water and air is nearly 10 μM) of its maximum capacity (24, 29). However, C3
plants do not concentrate CO2 using carbon concentrating mechanisms (CCMs)
probably because of having large amount of RuBisCO and the presence of highly
active β type of CA in the thylakoid stroma (Table 12.1). RuBisCO has both
oxygenase and carboxylase activity. Km for oxygenase activity is 700 times lower
than that of carboxylase activity. Therefore, increase in atmospheric CO2 concen-
tration is associated with not only the increase in CO2 assimilation but also
the reduction in photorespiration consuming significant amount of oxygen
[24]. However, at higher CO2 concentration, rate of photosynthesis is limited by
the regeneration of the ribulose-1,5-bisphosphate (RuBP). Triose phosphate
produced during Calvin cycle releases inorganic phosphate Pi which is essential
for ATP synthesis and RuBP regeneration from phosphorylated intermediates.
Therefore, under these circumstances, photosynthesis is called Pi or triose phos-
phate limited use limited [24]. Fixed CO2 can have three fates: (a) some of the CO2
releases back into the atmosphere through respiration; (b) some of the CO2 transfers
to the soil through root exudates, root death, litter fall, and coarse woody debris
which after decomposition releases back the CO2 into the atmosphere; and (c) some
of the CO2 gets stored in the wood [21]. Contrary to C3 plants, C4 plants have the
ability to grow efficiently in stress condition such as high temperature, low water
availability, high irradiance, and saline soils. C4 plants on land have thick-ended
walls of bundle-sheath cells, which help them in preventing the leakage of CO2
outside [30]. CCM activates in plants by spatial separation of CO2 fixation by
Table 12.1 Comparative study of CCMs occurring in plants (C3 and C4), microalgae, and cyanobacteria [29, 47]
HCO3
Km (CO2) accumulation
Type of CCM efficiency in of over ambient Carbon concentrating RuBisCO
photosynthesis DIC acquisition RuBisCO concentration DIC preference Location of CA mechanisms location
Plants (C3) Least efficient 15–25 μM N.A CO2 as substrate Highly active CA One carboxylation reaction to Chloroplast
in stroma of form C3 compound stroma of
chloroplast most of the
leaf cells
Plants (C4) Intermediately 28–60 μM N.A PEP carboxylase (affin- Abundance of CA Two carboxylation reactions to Bundle-sheath
efficient ity with HCO3 only) in the cytosol form C4 compound. Car- cells
more abundant than of mesophyll boxylation and decarboxyl-
C3 plants cells ation reactions are spatially
Absent in bundle- separated
sheath cells
Microalgae Less efficient than 20 μM 20 fold CO2 is preferred. Both Periplasmic HCO3/Na+ symporter; Na+/H+ Pyrenoid
cyanobacteria CO2 and HCO3 space; cyto- antiporter; NADHdh- inside
inside the cytosol sol, thylakoid induced transport; chloroplast
lumen; ATP-fueled HCO3 trans-
pyrenoid port; CO2 transporter
Cyanobacteria Most efficient 200 μM 100 fold Preference is less spe- Carboxysomes Direct transfer of HCO3; Carboxysomes
cific. Only HCO3 CA-mediated CO2 diffusion
inside cytosol
12 Carbon Dioxide Sequestration by Biological Processes 309

Calvin cycle. Normally PEP carboxylase catalyzes the CO2 fixation in the mesophyll
cells. C4 acids transported to bundle-sheath cells are decomposed to gaseous CO2 by
the action of NADP-malic enzyme and other decarboxylases. Both these processes
are temporarily separated in CAM plants. Detailed discussions on Calvin cycle and
CCMs in C4 and CAM plants are discussed in another section.

12.2.1.2 Oceanic Fertilization

Oceanic fertilization can be defined by adopting a practice of supplementing limiting


nutrients to the phytoplanktons, causing increase in photosynthetic efficiency and
CO2 removal rate. For example, micronutrient Fe was added in 10  10 km patches
of ocean, and 30 times increase in chlorophyll concentration with increase in
100,000 kg of carbon fixation was found [4]. However, its practice should be adopted
cautiously as it can negatively interfere with the marine ecosystem. In addition,
sinking organic matters on decomposition may produce other stronger greenhouse
gases such as methane and NOx [31].

12.2.1.3 Forest Fertilization

The study conducted by Oren et al. (2001) revealed that the enriched CO2
concentration (550 ppmv) had only marginal positive effect on the biomass carbon
increment. However, synergistic gain was observed when plants were grown in
enriched CO2 as well as nutrient condition. The gain was threefold and twofold
larger at the poor site and at the moderate site, respectively. Fertilizing at higher
CO2 concentration is more prominent than the ambient CO2 concentration [32].
Further, application of fertilizer in paddy soils has been found favoring the growth
of autotrophic microorganisms, resulting in increased CO2 sequestration [33].

12.2.2 CO2 Sequestration by Microalgae and Cyanobacteria

12.2.2.1 Microalgae

Photosynthesis in microalgae takes place in chloroplast. Apart from other pigments


such as β-carotene and xanthophylls, they have chlorophyll a and chlorophyll b as
the major pigments, which give them bright green color. Light requirement of the
typical algae is lower than the higher plants [34]. Light conversion efficiency and
productivity are proportional to the increase in light intensity till it attains saturation
light intensity [35]. Saturation light intensity of algae such as Chlorella and
Scenedesmus sp. is of order 200 μ mol m2 s1 [36]. Photosynthesis is coupled
with release of O2 as by-product. High dissolved oxygen (DO) in the culture
(>35 mg L1) has inhibitory effect on the photosynthesis [7, 37]. In the closed
310 K. Kumar and D. Das

photobioreactor, DO level has been found to increase as high as 400 %, thus


severally affecting the photosynthetic process [38]. Further, cations dissolved in
medium also negatively affect the photosynthetic process. Algae carry net negative
charge in their surface and hence are a potential adsorber of polyvalent cations
present in the surrounding medium. Adsorption of polyvalent cations on the surface
of algae causes morphological changes in the cell morphology or can replace/block
the prosthetic metal atoms in the active site of relevant enzymes leading to
photosynthesis inhibition [34].

12.2.2.2 Cyanobacteria

Cyanobacteria is a gram-negative bacteria. Compared to other gram-negative


bacteria, the cell wall of cyanobacteria has thicker peptidoglycan layer
[39]. Cyanobacteria lack organelles in the cell. Respiratory chain is located in
thylakoid membrane and plasma membrane. Photosynthesis takes place in thyla-
koid located in the cytoplasm, and photosynthetic electron transport takes place in
its membranes. The photosynthetic apparatus of cyanobacteria is similar to the
chloroplast of green algae and plants. However, major difference is in the antenna
system. Cyanobacteria lack chlorophyll b and depend primarily on chlorophyll a
and specialized protein complex known as phycobilisomes for harvesting light
energy. Light harvesting complex can be lipophilic as well as hydrophilic in nature.
Lipophilic pigments such as chlorophyll a and carotenoids are located within the
thylakoid membranes whereas hydrophilic antenna pigments are housed in
phycobilisomes attached outside of thylakoid membranes. Examples of hydrophilic
pigments are allophycocyanin (APC), phycocyanin (PC), phycoerythrin (PE), or
phycoerythrocyanin (PEC) [40]. Similar to green algae and plants, cyanobacteria
carry out oxygenic photosynthesis, releasing O2. However, in anaerobic condition,
they can also carry out anoxygenic photosynthesis using PSI. Sources of electrons
are smaller organic compounds such as succinate, hydrogen sulfide, and thiosulfate
instead of water [41]. Anoxygenic photosynthesis results in generation of ATP by
cyclic photophosphorylation around the PSI.

12.2.2.3 Photosynthesis: Key for CO2 Sequestration

Photosynthesis, the key biological process occurring in plants and a wide number of
microorganisms such as algae and cyanobacteria, helps in mitigating atmospheric CO2.
Photosynthesis occurs in the chloroplasts of plants and algae. Chloroplasts contain the
thylakoid vesicles arranged in stacks, containing photosynthetic apparatus. Contrary to
plants and algae, prokaryotic cyanobacteria do not have fixed organelles for keeping
photosynthetic apparatus. They are located in cytoplasm as free and isolated photo-
synthetic lamellae [42]. Biological system uses photosynthesis for harnessing solar
energy for the preparation of food with the help of CO2 and water. Photosynthesis
consists of two distinct processes: light-dependent process and light-independent
12 Carbon Dioxide Sequestration by Biological Processes 311

Fig. 12.1 Schematic diagram of Z scheme of photosynthetic process involving PSII and PSI. Solid
and dashed lines show the noncyclic and cyclic flow of electrons. Abbreviation: PQA plastoquinone,
PQB second quinine, Chl ao chlorophyll a, Chl a1 phylloquinone, PC plastocyanin [53]

process. In light-dependent process, they conserve light energy in the form of ATP and
NADPH with the help of chlorophyll and light harvesting complex. Light-independent
process utilizes ATP and NADPH for the fixation of CO2 into triose phosphates, starch,
sucrose, and other derived products. Photosynthetic apparatus of plants, algae, and
cyanobacteria are similar in nature. Photosynthetic apparatus are found in the thylakoid
membranes consisting of protein complex, electron carriers, and lipid molecules. The
electron carriers are arranged in the shape of 90 tilted English alphabet Z. Two
reaction centers, PSII and PSI, help in exciting the electrons by absorbing the light of
680 nm and 700 nm, respectively (Fig. 12.1). Two reaction centers are connected with
series of electron carriers such as plastoquinone, cytochrome b6f complex, and
plastocyanine arranged in fixed increasing order of redox potential. Electron carriers
between two centers help in transferring excited electrons at PSII to pass to PSI
smoothly. Electrons at PSII come from breaking of water molecules into oxygen
molecules, protons, and electrons. Two moles of water dissociates into four moles of
protons and electrons and one mole of oxygen molecule (Eq. 12.1) [42].

2H2 O ! 4Hþ þ 4e þ O2 ð12:1Þ

Protons accumulate into the lumen of thylakoid membranes, generating proton


gradient across the membrane. Concentration gradient drives protons outside thy-
lakoid membrane through ATP synthase producing ATP. Thus, ATP synthase
312 K. Kumar and D. Das

produces not only energy for the cell but also avails protons for the reduction of
NADP+ to produce NADPH. At PSII, light energy causes charge separation
between P680 and pheophytin, creating P680+/Pheo. Pheophytin transfers the
electrons to a permanently bounded molecule (QA) to photosystem II. At QA site
plastoquinone receives single electron instead of two. QB and QA sites differ from
each other as the former requires two electrons to reduce instead of one, which is the
case of the latter, causing two turnovers for the complete reduction of plastoquinone
at QB. Due to close proximity of QB site, protons added to plastoquinone during its
reduction come from the outside aqueous phase of the membrane [13]. At
photosystem I, most of the antenna chlorophyll molecules are attached to the
reaction center proteins [13]. Plastoquinone then transfers their electrons to next
electron carrier cytochrome b6f complex with simultaneous release of protons into
the lumen. Cytochrome b6f complex is attached with membrane-bound protein
complex. Plastocyanin (PC) acts as a last protein carrier delivering electrons to
PSI. PSI again excites the electrons received from electron carriers. Here, excited
electrons are used to reduce ferredoxin (Fd), a protein loosely attached with
thylakoid membranes from outside. Reduced Fd interacts with ferredoxin NADP+
reductase (FNR) and the latter catalyzes the reduction of NADP+ to NADPH as
shown in (Eq. 12.2).

2Fdred þ 2Hþ þ NADPþ ! 2Fdox þ NADðPÞH þ Hþ ð12:2Þ

Each of the photosystem contributes one photon to the transfer of one electron.
Therefore, two photons are required for the transfer of one electron along the
electron carriers to the NADP+. Two moles of water and eight photons are required
to produce two moles of NAD(P)H as shown in (Eq. 12.3).

2H2 O þ 2 NADPþ þ 8 photons ! O2 þ 2NADðPÞH þ 2Hþ ð12:3Þ

Reduction of one molecule of CO2 requires two moles of NAD(P)H. Energy


content in photon is 480 KJ/mol, whereas energy gathered after reduction of one
reduced CO2 molecule is equivalent to 1,400 KJ. Therefore, theoretical maximum
efficiency of the process is 36.4 %.

12.2.2.4 Carbon Concentrating Mechanisms (CCMs)

Cyanobacteria and algae behave like C3 plants but have much less affinity for CO2
[29]. However, they develop CCMs similar to C4 and CAM plants, causing better
photosynthetic efficiency compared to C3 plants [30]. Their RuBisCO have
oxygenase as well as carboxylase activities, depending upon the concentration of
local CO2 in the compartment housing RuBisCO. Carboxylase activity of RuBisCO
activates under high CO2 pressure, whereas low concentration of CO2 forces it to
undergo oxygenase activity. Oxygenase reaction has many disadvantages. Firstly,
glycolate 2-phosphate is the end product of oxygenase activity of RuBisCO
12 Carbon Dioxide Sequestration by Biological Processes 313

(Eq. 12.4). It has no use to algal cells; therefore, significant amount of cellular
energy is wasted by consuming it. Secondly, it releases previously fixed CO2,
during the carboxylase activity of RuBisCO, which causes the loss of nearly
50 % of algal biomass [43]. Metabolism of glycolate 2-phosphate produces glycine,
which on condensing with another glycine molecule produces serine, resulting in
the loss of CO2 [44]. Thirdly, loss of fixed carbon further negatively interferes with
the regeneration of RuBP, required for the smooth functioning of the cycle.

Ribulose 1, 5 bisphosphate þ O2 ! glycerate 3-phosphate


þ glycolate 2-phosphate ð12:4Þ

Affinity constant (Km) of microalgae and cyanobacteria for CO2 is high com-
pared to C3 plants, indicating low affinity of RuBisCO for CO2 (Table 12.1). Under
normal atmospheric air, RuBisCO is only half-saturated with CO2 [7, 29]. The
diffusion of CO2 in aqueous solution is 10,000 times slower than the CO2 diffusion
in air [30]. Poor diffusion of atmospheric CO2 in aqueous solution and low affinity
of CO2 by algae, cyanobacteria, and some chemoautotrophic bacteria pose a
limitation to the carboxylase activity of RuBisCO [45]. Therefore, for maintaining
this, it is necessary to saturate the compartment containing RuBisCO with CO2.
Fortunately, algae have carbon concentrating mechanisms (CCMs), which help
them to increase the local CO2 even when there is lower CO2 concentration outside
the algal cells [46]. Contrary to plants, algae and cyanobacteria have single-cell
CO2 concentrating mechanisms. However, they have internal compartments within
the chloroplasts (pyrenoid in algae and carboxysome in cyanobacteria). Increase in
concentration of CO2 at close proximity of RuBisCO has many advantages. Firstly,
it activates RuBisCO. Secondly, it increases carboxylase activity of RuBisCO, and,
thirdly, dissolved inorganic carbon (DIC) influx may help in maintaining internal
pH and dissipating excess light energy [47]. CCMs can be divided broadly into two
categories: biochemical and biophysical CO2 pumps. CCMs help in accumulation
of large number of intracellular inorganic carbon inside the RuBisCO compartment.
However, inorganic carbons are not available for fixation by RuBisCO. They must
be converted back into gaseous CO2 for the action of RuBisCO. CA is a vital
enzyme, which overexpresses under low external CO2 environment. CA acts as a
catalyst in the interconversion of inorganic carbon back to gaseous CO2.

Biochemical CO2 Pump

C4 Mechanism – Biochemical pumps are mostly found in the terrestrial plants.


However, some phytoplanktons and macroalgae also have evidence of biochemical
pumps. C4 mechanisms have also been found in macroalgae such as Udotea
flabellum and Thalassiosira weissflogii [43]. The role of C4 mechanism is to
biochemically transport the DIC from the site excessive in it to the site where
RuBisCO is active [43]. As the name indicates, CO2 is stored in a four-carbon
314 K. Kumar and D. Das

compound, oxaloacetate. Phosphoenolpyruvate (PEP) is the carrier molecule,


which combines with HCO3 using PEP carboxylase to form oxaloacetate. PEP
carboxylase utilizes bicarbonates rather than CO2; therefore, the gaseous CO2
entering into the mesophyll cell must be rapidly converted to bicarbonate with
the help of carbonic anhydrase [48]. At physiological CO2 levels and pH, Km
(HCO3) of PEP carboxylase and HCO3 concentration in the cytoplasm of
mesophyll cells were estimated to be about 8 μM and 50 μM, respectively. There-
fore, unlike RuBisCO, PEP carboxylase is always saturated with HCO3 at ambient
CO2 concentration. Therefore, CA is mostly confined to mesophyll cell in the C4
plants compared to chloroplast in C3 plants [49]. Oxaloacetate further readily
reduced to malate by the action of NADP-malate dehydrogenase. In plants, malate
is transported to bundle-sheath cell where it is decarboxylated by the action of
NADP-malic enzyme. At the site of RuBisCO, malate can be decarboxylated to
pyruvate and releases gaseous CO2 for the action of RuBisCO enzyme in Calvin
cycle.
CAM Mechanism – CAM plants are usually found in desert area. In cool night,
guard cells open to receive CO2, while in daytime, it is closed to prevent water loss.
This mechanism is also proposed in brown macroalgae for the assimilation of
photosynthetic inorganic carbon [43]. PEP comes from the starch accumulated
during daytime using Calvin cycle. Enzymes and compounds taking part in CAM
mechanism are similar to C4 mechanism. However, end storage compound malate
is temporarily separated over time rather than spatially as in C4 plants [43]. PEP
carboxylase catalyzes the reaction of PEP and HCO3 to form oxaloacetate.
NADP-malic dehydrogenase reduces oxaloacetate to malate, which is transported
to vacuole having low pH at night. In daytime, whole pathways get reversed back
and malic acid transported back to cytosol for the decarboxylation reaction,
flooding cytosol with CO2. Guard cell is closed during the daytime, preventing
CO2 diffusion outside the cell. It is to be noted that PEP carboxylase activity is also
under control to prevent wasteful synthesis of C4 compounds during daytime.

Biophysical CO2 Pump

CCMs in Cyanobacteria
Being simple in structure, algae and cyanobacteria cannot stop the diffusion of CO2
[30]. Carboxysome is the key for the success of CCMs in cyanobacteria. It acts as a
storehouse for CO2 having limited permeability for CO2 leakage. The ultimate
target of CCMs is to transport DIC to the storehouse, carboxysome. Along with
cyanobacteria, carboxysomes are also found in some chemoautotrophic bacteria
growing in CO2 concentration lower than the Km of RuBisCO [45]. Carboxysomes
have mainly icosahedral structures with the diameter of 100–200 nm. The number
of carboxysomes present in per cell of cyanobacteria varies from 5 to 20 depending
upon the growth conditions and species to species [45].
12 Carbon Dioxide Sequestration by Biological Processes 315

Fig. 12.2 Schematic diagram of CCMs occurring in cyanobacteria [29, 43, 49]

Most of algae and cyanobacteria examined so far have been found to have
transporter for both CO2 and HCO3. However, few algae can also assimilate either
CO2 or HCO3. CCMs are extensively studied in cyanobacteria compared to
microalgae. Active transport occurs across the membrane having low permeability
to DIC. In cyanobacteria, transport of CO2 or HCO3 occurs via plasmalemma or
thylakoid membranes. All the DIC removed from the outside by various DIC
transporters are delivered only in the form of HCO3 into the cytosol of cell
[45]. Bicarbonate concentrating ability of cyanobacteria is nearly five times greater
than the microalgae, causing higher photosynthetic and CO2 consumption
efficiency of the former [30]. Outside, DIC including CO2 penetrate the plasma
membrane and reach the closest membrane of the chloroplast with the help of
various transporter proteins. At least three mechanisms of active transport have
been proposed in cyanobacteria (Fig. 12.2) [29, 49, 50]: (1) HCO3 can be
transported to cytosol with the help of ABC-type transporter utilizing energy in
the form of ATP, which is a high-affinity, low-CO2-induced, and sodium-
independent mode of HCO3 assimilation; (2) HCO3 transport into cytosol may
also be the result of HCO3/Na+ symporter or the regulation of pH through Na+/H+
antiporter; and (3) for the active transport of CO2, NADH dehydrogenase may have
constitutively low or inducible high affinity for CO2. In either case, CO2 is first
converted to HCO3 by NADHdh at plasmalemma of cyanobacteria and then
316 K. Kumar and D. Das

transported into the cytosol. Intracellular pH of cyanobacteria is near to 8, causing


HCO3 as predominant species inside the cytoplasm (equilibrium ratio of HCO3
and CO2 is near to 100). HCO3 is the charged ion and therefore cannot escape the
lipid bilayer of the cyanobacteria [45]. HCO3 then diffuses into carboxysome
where it is acted upon by CA enzyme to flood the compartment with gaseous CO2
for the action of RuBisCO of Calvin cycle [43]. Remaining amount of CO2 in
carboxysome is diffused outward [47]. In addition, it has also been proposed to have
direct access of HCO3 to cytoplasm of cyanobacteria. Rate of direct transport of
HCO3 across plasma membrane is less significant compared to its active transport
[51]. In yet another mechanism, compartment containing high concentration of
HCO3 is acidified using proton pump and flooding with CO2 (Eq. 12.5). In acidic
compartment, HCO3 decomposes into CO2 by the action of CA or proton-driven
catalysis of HCO3 to CO2. The high level of CO2 then diffuses into the more
alkaline compartment housing RuBisCO for the action of RuBisCO [43].

Alkaline ðHCO3  Þ ! Acidic ðHCO3  ! CO2 Þ ! Alkaline ðCO2 Þ ð12:5Þ

CCMs in Algae
Compared to cyanobacteria, CCMs in microalgae are less understood because of
more compartments inside the cell and very diverse group of microorganisms
[29]. Microalgae composed of one or few cells do not have impermeable cell
walls like plants to prevent leakage of CO2. The challenging task is to prevent the
leakage of concentrated CO2 while allowing other nutrients to come in. The
efficiency of acquisition of DIC depends on environmental conditions. For exam-
ple, acquisition of DIC in low atmospheric CO2 was found higher compared to high
atmospheric CO2 condition. However, the amount of RuBisCO did not change
during adaptation from low to high CO2 condition. This indicated the existence of
transport system for the uptake of DIC into the cells. Cyanobacteria and microalgae
can accumulate nearly 100- and 20-fold increase in HCO3 within the cells,
respectively, over ambient CO2 level [29]. DIC accumulation in algae is generally
lower than cyanobacteria probably due to higher affinity of RuBisCO present in the
former for CO2. CCMs in microalgae have been hypothesized similar in nature as
cyanobacteria by most of the researchers. Similar to carboxysome of cyanobacteria,
microalgae also have a compartment (pyrenoid) in chloroplast densely packed with
RuBisCO. Accumulation of charged HCO3 ions lessens the chance of leakage.
Freshwater and marine green algae are capable of utilizing HCO3. However, in
most of the microalgae, CO2 is main form of carbon entering into the cell and
HCO3 in chloroplast. In Chlamydomonas reinhardtii, active transport of CO2 has
been found to have preference over HCO3 [51]. CO2 uptake in whole cell is due to
diffusion, while through the chloroplast, it is mediated by transfer [47]. A range of
CAs participate in each of the compartment to maintain the equilibrium between
CO2 and HCO3 (Fig. 12.3). Eventually DIC entering into the pyrenoid is in the
form of HCO3, which needs to be converted back to CO2 to enrich RuBisCO
compartment for the carboxylase reaction. Cell membrane of pyrenoid does not
12 Carbon Dioxide Sequestration by Biological Processes 317

Fig. 12.3 Schematic diagram of CCMs occurring in green algae [29, 43, 49]

allow CO2 to leak out, allowing sufficient time for RuBisCO to use CO2 in the
Calvin cycle. CA is an important enzyme for the successful operation of CCM. CA
in microalgae is of different type and located in different places inside the cell. It
has been found in periplasmic space, cytosol, as well as inside the pyrenoid. Direct
uptake of HCO3 and CO2 diffusion facilitated by periplasmic CA across the
plasma membrane have been proposed [29]. Periplasmic CA probably helps in
the diffusion of CO2 and supply of HCO3 into the cytosol across the plasma
membrane. However, in every case examined so far, electrochemical potential
gradient for HCO3 was outward. By examining the HCO3 transport in Ulva,
HCO3 uptake in microalgae has been proposed to occur through the HCO3/OH
antiport [47].

12.2.3 Bacteria

The non-photosynthetic bacteria play an important role in global carbon cycle.


Advantages of these microorganisms are their ability to adapt to and survive
extreme conditions of the environment. However, consumption of large amount
of H2 as electron donor limits their practical application [52]. Enzymes involved in
318 K. Kumar and D. Das

the CO2 fixation pathway are sensitive to oxygen; therefore, they grow best in
anaerobic environment [52]. Their ability to sequester CO2 under anaerobic
environment can be helpful for their application in O2-deficient but CO2-rich
environment such as under the soil and flue gas [52].
Photosystem of some of the bacteria looks like either PSI or PSII. Lack of PSII
deprives bacteria to use electrons of water and evolve oxygen as by-product of
photosynthesis. Some simple inorganic or organic molecules substitute the water for
the electrons needed to reduce CO2 into useful simple sugar. They have bacteriochlo-
rophyll, a family of molecules similar to chlorophyll, but absorb light in the range of
700–1,100 nm. Similar to oxygenic photosynthesis, electron transfer results in the
generation of proton gradient across the thylakoid membrane. Outward protein gradient
drives the proton out through ATP synthase, causing synthesis of ATP. Energy for CO2
reduction comes from ATP and NADH. Electron carriers are quinone such as
ubiquinone, menaquinone, and the cytochrome bc complex, which work similarly to
cytochrome b6f complex present in the oxygenic photosynthetic system [13].
Green gliding bacteria such as Chloroflexus aurantiacus harvest light using
chlorosomes similar to green sulfur bacteria. CO2 fixation in these microorganisms
does not involve Calvin or Krebs cycle. They usually do photosynthesis under
anaerobic condition. Green and purple bacterial membranes are in the form of
lamellae, vesicles, or specialized structures (chlorosomes) where light reaction
takes place. Sulfur purple bacteria such as Chromatium vinosum fix CO2 for their
survival using Calvin cycle.

12.2.3.1 Purple Bacteria

Photosynthetic machinery of purple bacteria such as Rhodospirillum rubrum,


Rhodopseudomonas viridis, and Rhodobacter sphaeroides is of pheophytin-quinone
type. Pheophytin is similar to chlorophyll but lacks central Mg2+ ions. Purple bacteria
have single reaction center called P870. Electrons at the reaction center get excited,
absorbing light of 870 nm. Excited electrons pass to cytochrome bc1 complex
through pheophytin and quinine sequentially [53, 54]. Cytochrome bc1 complex is
the hub that pumps the protons to generate proton gradient and electrons back to
reaction center via cytochrome c2 (Fig. 12.4a). Light-driven cyclic flow of electrons
enables to produce ATP using ATP synthase. Purple bacteria are of two types:
non-sulfur purple, such as Rhodopseudomonas viridis and Rhodobacter sphaeroides,
and sulfur purple bacteria, such as Chromatium vinosum. Electron source for
non-sulfur bacteria is organic compounds such as malate and succinate, whereas
sulfur purple bacteria extract electrons from inorganic sulfur compounds such as
hydrogen sulfide. CO2 is fixed in purple bacteria using Calvin cycle.

12.2.3.2 Green Sulfur Bacteria

Reaction center of green sulfur bacteria such as Chlorobium vibrioforme and


Chlorobium thiosulfatophilum is similar to PSI of oxygenic photosynthesis.
12 Carbon Dioxide Sequestration by Biological Processes 319

Fig. 12.4 Schematic diagram of photosynthetic process involving only PSI in (a) purple bacteria
and (b) green sulfur bacteria. Solid and dashed lines show the noncyclic and cyclic flow of
electrons. Abbreviation: Q quinine [53]

Similar to phycobilisomes of cyanobacteria, the antenna system of the green


sulfur bacteria such as bacteriochlorophyll and carotenoids is contained in
complexes known as chlorosomes attached to the surface of the photosynthetic
membrane through baseplate containing antenna bacteriochlorophyll a [13].
Reaction center involved in green sulfur bacteria is called Fe–S type. Therefore,
green sulfur bacteria are not dependent on the reverse electron flow for the carbon
reduction as the reduced ferredoxin uses its electrons to reduce NAD+/NADP+
using ferredoxin–NAD(P)+ oxidoreductase enzyme. At reaction center, electrons
get excited by absorbing light intensity of 840 nm. Excited electrons can follow
two pathways: one cyclic flow of electrons back to the reaction center via Q,
Cyt bc1 complex, and Cyt c553, and another noncyclic flow of electrons through
iron–sulfur protein ferredoxin (Fd) to reduce NAD+ to NADPH using ferredoxin/
NAD reductase (Fig. 12.4b) [53, 54]. Similar to purple bacteria, protons are
pumped by the cytochrome bc1 complex to generate ATP. Electrons at reaction
center are replaced by the oxidation of H2S to elemental So and then to
SO42. Electron carriers in green sulfur bacteria are much better placed according
to their electronegativity than the purple bacteria which ensure reduction of NAD
without the need of reverse electron flow. Green sulfur bacteria reduce free CO2
by reversing original tricarboxylic acid cycle (Krebs cycle) with the input of
energy. Thus, green sulfur bacteria can fix CO2 even in the absence of RuBisCO.
Green sulfur bacteria extract both electrons and hydrogen from sulfur
compounds [13].
320 K. Kumar and D. Das

12.3 CO2 Sequestration Pathways

12.3.1 Calvin Cycle

CO2 is fixed into carbohydrate in light-independent stage using Calvin–Benson cycle.


Proteins participating in CO2 fixation have been found outside the thylakoid mem-
brane in aqueous phase [13]. CO2 fixation reaction is catalyzed by the carboxylase
activity of ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO) [55]. Calvin
cycle can be divided into carboxylation, reduction, and regeneration reaction
(Fig. 12.5). In carboxylation reaction, three molecules of CO2 combine with three
molecules of ribulose-1,5-bisphosphate (5 C) using carboxylase activity of RuBisCO
enzyme to form six molecules of glycerate 3-phosphate (Eq. 12.6). Glycerate
3-phosphate further reduces to glyceraldehyde 3-phosphate in the reduction step.
Reduction reaction is followed by regeneration reaction where ribulose-1,5-
bisphosphate, the starting material of Calvin cycle, is regenerated using five molecules
of glyceraldehyde 3-phosphate, while the remaining one molecule of glyceraldehyde
3-phosphate is used for the synthesizing biosynthetic material [56].

Ribulose 1, 5 bisphosphate þ CO2 þ H2 O ! 2 glyceraldehyde 3-phosphate ð12:6Þ

Fig. 12.5 Schematic diagram of Calvin cycle


12 Carbon Dioxide Sequestration by Biological Processes 321

Standard free energy for the synthesis of one mole of glucose is equal to
2,870 KJ [13]. Photosynthetic process is further generalized by Van Niel (Eq. 12.7).

CO2 þ 2H2 A þ light ! ðCH2 OÞn þ 2A þ H2 O ð12:7Þ

where A is O and S for oxygenic photosynthesis and anoxygenic photosynthesis


(taking H2S as electron donor), respectively. Later, it was demonstrated by Van Niel
that molecular oxygen comes from dissociation of water rather than CO2. The end
product of Calvin cycle, glyceraldehyde 3-phosphate, is used in synthesizing cellular
biosynthetic material for immediate energy source and sucrose that is transported to
cytosol for storage as starch in chloroplast of green algae and glycogen in
cyanobacteria [57]. Considering acetyl-CoA as the end product, seven molecules of
ATP and four molecules of NAD(P)H are required to reduce two molecules of CO2.
Thus, Calvin cycle is the most energy intensive pathway for CO2 sequestration [58].

2 CO2 þ 7 ATP þ 4 NADðPÞH þ HS-CoA ! Acetyl-CoA þ 7 ADP


þ 4 NAD ðPÞþ ð12:8Þ

12.3.2 Wood–Ljungdahl (WL) or Reductive


Acetyl-CoA Pathway

Wood–Ljungdahl (WL) or reductive acetyl-CoA pathway is a bidirectional pathway,


which is predominant primarily in strict anaerobic bacteria and archaea of phyla
Firmicutes and Euryarchaeota, respectively [58]. This pathway is used for chemoau-
totrophic carbon fixation in acetogens such as Clostridium thermoaceticum and
Acetobacterium woodii and methanogens such as Methanobacterium thermoauto-
trophicum and most autotrophic sulfate reducers such as Desulfobacterium
autotrophicum. Wood–Ljungdahl (WL) pathway was found to be the most efficient
non-photosynthetic pathway based on the most expensive substrate (i.e., H2 or elec-
trons), for the production of acetate and ethanol [58]. Reductive acetyl-CoA pathway
uses only four moles of H2 to form one mole of acetate. Two molecules of CO2 directly
combine to form one molecule of acetate. This pathway is different from other six
known CO2 fixation pathways as it does not undergo in cyclic manner. In addition to
the use of Fd for reduction reaction, carbon monoxide dehydrogenase and acetyl-CoA
synthase are the main enzymes involved in this pathway. These enzymes are very
sensitive to oxygen. It has both carbonyl as well as methyl components (Fig. 12.6). Out
of two molecules of CO2, one is reduced to carbonyl group (C¼O) catalyzed by CO
dehydrogenase, and the other CO2 is captured on special tetrahydrofolate cofactor and
reduced to a methyl group. Carbonyl group bonded with enzyme is combined with the
methyl group to form acetyl-CoA by enzyme acetyl-CoA synthase complex. The
reducing equivalents for the pathway are obtained by oxidation of molecular hydrogen
during autotrophic growth or NADH and reduced ferredoxin during heterotrophic
growth [59]. For the formation of one molecule of acetyl-CoA using two molecules
322 K. Kumar and D. Das

Fig. 12.6 Schematic


diagram of
Wood–Ljungdahl (WL) or
reductive acetyl-CoA
pathway. (Reprinted from
Ref. [60]. Copyright 2011,
with permission from
Elsevier)

of CO2, WL pathways need one molecule of ATP and four molecules of NAD(P)H
(Eq. 12.9) [58]. Among the four pathways discussed above, WL pathway is most
efficient based on CO2 sequestration per ATP consumed. Moreover, ATP consumed in
WL pathway is even less than one, as some of the energy is conserved in the membrane
gradient in the form of ATP. rTCA cycle comes next to WL pathway in terms of energy
efficiency. Both the pathways are found in highly reduced environments in which
carbon reduction is more favorable [58].

2 CO2 þ 1 ATP þ 4 NADðPÞH þ HS-CoA ! Acetyl-CoA þ 1 ADP


þ 4 NAD ðPÞþ ð12:9Þ

12.3.3 Reductive Tricarboxylic Acid Cycle (rTCA)


or Reverse Citric Acid Cycle

Reductive tricarboxylic acid cycle (rTCA) has been found in bacteria under
anaerobic or microaerobic conditions. Bacteria dwelling under these conditions
are green sulfur bacteria such as Chlorobium sp. (Chlorobium limicola, Chlorobium
12 Carbon Dioxide Sequestration by Biological Processes 323

Fig. 12.7 Schematic diagram of (a) reductive tricarboxylic acid cycle (rTCA) or reverse citric
acid cycle shown in green arrow and (b) dicarboxylate/4-hydroxybutyrate cycle shown in red
arrow. Black arrows showing reaction pathways are common to both of them (Reprinted from
Ref. [60]. Copyright 2011, with permission from Elsevier)

thiosulfatophilum, etc.), sulfur-reducing bacteria (Desulfobacter), knallgas bacteria


or hydrogen-oxidizing bacteria (Hydrogenobacter, Aquifex), and archaea
(Thermoproteus). They utilize this cycle to fix atmospheric CO2 through
anoxygenic photosynthesis [59]. Reductive citric acid cycle is oxidative citric
acid cycle running in reverse direction [61]. In TCA cycle, one molecule of
acetyl-CoA breaks down to form two molecules of CO2 and energy. However, in
reverse TCA cycle, two molecules of CO2 are used to synthesize one molecule of
acetyl-CoA using 2 ATP and H+ from NADH and NADPH (Fig. 12.7a). Two
carbon fixing enzymes, pyruvate synthase (pyruvate:ferredoxin oxidoreductase)
and 2-oxoglutarate synthase (2-oxoglutarate:ferredoxin oxidoreductase), are the
key enzymes of rTCA cycle and were found to be dependent on Fd [61, 62]. Two
reactions involving ferredoxin (Fd) significantly help in reversal of some of the
reactions, which are otherwise nonreversible in nature. Another enzyme citrate
lyase was discovered which is dependent on ATP [63]. ATP-citrate lyase cleaves
citrate, a six-carbon compound, into oxaloacetate, a four-carbon compound, and
acetyl-CoA. These three enzymes work together to make energetically unfavorable
reverse reactions possible. One molecule of succinyl-CoA combines with one
molecule of CO2 to form one molecule of 2-oxoglutarate using 2-oxoglutarate
synthase. 2-Oxoglutarate further takes one molecule of CO2 to form isocitrate
using isocitrate dehydrogenase (IDH). Aconitase converts isocitrate to citrate,
which is acted upon by citrate lyase to form oxaloacetate and acetyl-CoA.
Acetyl-CoA combines with another molecule of CO2 to form pyruvate using
pyruvate synthase. It is followed by synthesis of PEP which combines with another
molecule of CO2 to regenerate oxaloacetate or other intermediates of the cycle in an
324 K. Kumar and D. Das

Fig. 12.8 Schematic diagram of (a) 3-hydroxypropionate cycle shown in light blue arrow, (b)
3-hydroxypropionate/4-hydroxybutyrate cycle shown in purple arrow. Black arrows showing
reaction pathways are common to both of them (Reprinted from Ref. [60]. Copyright 2011, with
permission from Elsevier)

anaplerotic manner [61]. Shiba et al. (1985) reported direct conversion of pyruvate
into oxaloacetate in H. thermophilus by an enzyme pyruvate carboxylase [64].
In one complete cycle, four molecules of CO2 are fixed to generate one molecule
of oxaloacetate, which itself is an intermediate of the cycle. It requires two
molecules of CO2, 2 ATP, and 4 NAD(P)H for the formation of one molecule of
acetyl-CoA as the end product as shown in Eq. (12.10) [58].

2 CO2 þ 2 ATP þ 4 NADðPÞH þ HS-CoA ! Acetyl-CoA þ 2 ADP


þ 4 NADðPÞþ ð12:10Þ

12.3.4 3-Hydroxypropionate Cycle

CO2 fixation in phototrophic green non-sulfur bacteria takes place using


3-hydroxypropionate pathway unique to eubacterium Chloroflexus aurantiacus
and some chemotrophic archaebacteria such as Sulfolobus metallicus, Ignicoccus
hospitalis, Acidianus brierleyi, and Acidianus infernus [60, 65, 66]. Enzymes
participating in this pathway are not sensitive to oxygen [60, 67]. Acetyl-CoA is
the starting precursor of 3-hydroxypropionate cycle. It combines with HCO3 to
form malonyl-CoA with the help of ATP-dependent acetyl-CoA carboxylase
(Fig. 12.8a). Acetyl-CoA carboxylase is also an essential enzyme for the synthesis
of fatty acid. It is followed by the conversion of malonyl-CoA to
12 Carbon Dioxide Sequestration by Biological Processes 325

3-hydroxypropionate by malonyl-CoA reductase in an NADPH-dependent


reaction. Malonyl-CoA reductase is a bifunctional enzyme, which reduces
malonyl-CoA to 3-hydroxypropionate via malonate-semialdehyde as an intermedi-
ate using its aldehyde dehydrogenase and alcohol dehydrogenase domain. Reduc-
tive conversion of 3-hydroxypropionate to propionyl-CoA is catalyzed by
propionyl-CoA synthase. Propionyl-CoA synthase is a trifunctional enzyme and
formally requires three enzymatic reactions. In the first step, activation to
3-hydroxypropionyl-CoA is catalyzed by CoA ligase, which is followed by dehy-
dration of 3-hydroxypropionyl-CoA to acrylyl-CoA by an enoyl-CoA hydratase.
Finally, acrylyl-CoA is reduced to propionyl-CoA by an enoyl-CoA reductase using
NADPH [65, 68, 69]. Propionyl-CoA undergoes carboxylation, catalyzed by
ATP-dependent propionyl-CoA carboxylase to form methylmalonyl-CoA. Isomer-
ization of methylmalonyl-CoA takes place in two sequential steps, catalyzed by
methylmalonyl-CoA epimerase and methylmalonyl-CoA mutase to form succinyl-
CoA. Succinyl-CoA transfers CoA for malate activation and forms succinate and
malyl-CoA [65, 68, 69]. Malyl-CoA, a four-carbon compound, breaks down by
malyl-CoA lyase to regenerate starting acetyl-CoA and glyxolate. Carboxylations
of acetyl-CoA and propionyl-CoA are the main CO2 fixation reactions. It is to be
noted that actual substrate for both the carboxylation reactions is HCO3 rather than
CO2. Each turn of cycle results in net fixation of two molecules of bicarbonate to
produce one molecule of glyxolate. Glyoxylate is considered as the initial CO2
fixation product. Glyxolate is further utilized in the synthesis of the cellular
material. Intermediate 3-hydroxypropionate is the characteristic of the cycle,
which on reductive conversion produces propionyl-CoA [65, 68, 69]. In
3-hydroxypropionate cycle, six molecules of ATP and four molecules of NAD(P)
H are required to reduce two molecules of CO2 to form one molecule of acetyl-CoA
as shown in Eq. (12.11) [58].

2 CO2 þ 6 ATP þ 4 NADðPÞH þ HS-CoA ! Acetyl-CoA þ 6 ADP


þ 4 NAD ðPÞþ ð12:11Þ

12.3.5 Other Pathways

12.3.5.1 Dicarboxylate/4-Hydroxybutyrate Cycle

Dicarboxylate/4-hydroxybutyrate cycle is the newly discovered autotrophic


pathway for carbon dioxide fixation in Ignicoccus hospitalis (Desulfurococcales),
Thermoproteus neutrophilus, and an anaerobic member of Thermoproteales
[70–72]. Evidence of dicarboxylate/4-hydroxybutyrate cycle has been reported by
Ramos-vera et al. (2009) in T. neutrophilus [72]. This pathway can also be divided
into two parts: the first part involves formation of succinyl-CoA from acetyl-CoA
using two inorganic carbons and the second part deals with regeneration of
acetyl-CoA from succinyl-CoA. Thus, this pathway has similarity with both
326 K. Kumar and D. Das

reductive tricarboxylic acid cycle (rTCA) and 3-hydroxypropionate/


4-hydroxybutyrate cycles (Fig. 12.7b). The first part of the cycle involves the
intermediates of reductive tricarboxylic acid cycle (rTCA) and the second part
involves the intermediates of 3-hydroxypropionate/4-hydroxybutyrate cycle via the
route similar to 4-hydroxybutyrate pathway to succinyl-CoA using pyruvate
synthase and pyruvate carboxylase, as carboxylating enzyme. So, the only difference
between this and the 3-hydroxypropionate/4-hydroxybutyrate cycle is the way
succinyl-CoA is created.

12.3.5.2 3-Hydroxypropionate/4-Hydroxybutyrate Cycle

Another pathway has been reported in some bacteria called 3-hydroxypropionate/


4-hydroxybutyrate cycle. This pathway was found in aerobic autotrophic members
of Sulfolobales. 3-Hydroxypropionate/4-hydroxybutyrate pathway was discovered
in Metallosphaera sedula which was earlier believed to fix CO2 using
3-hydroxypropionate cycle. Malyl-CoA lyase, an enzyme used in the regeneration
of acetyl-CoA, was absent in the cell extract of M. sedula, resulting in the proposed
alternative pathway for the regeneration of starting material of cycle. Therefore,
3-hydroxypropionate and 3-hydroxypropionate/4-hydroxybutyrate cycle share the
same steps from acetyl-CoA to succinyl-CoA (Fig. 12.8b). However, the enzymes
involved in the reaction steps from acetyl-CoA to succinyl-CoA are not same in
both cases. Detailed structure of acetyl-CoA carboxylase participating in
3-hydroxypropionate cycle reveals that the enzyme is composed of four subunits.
Contrary to this, acetyl-CoA carboxylase taking part in 3-hydroxypropionate/
4-hydroxybutyrate cycle has only three subunits. Similarly, malonyl-CoA reductase
of both the cycles is not the same. They differ in the different intermediates through
which regeneration of acetyl-CoA takes place from succinyl-CoA. The intermediates
between acetyl-CoA and succinyl-CoA in 3-hydroxypropionate/4-hydroxybutyrate
cycle are succinic semialdehyde, 4-hydroxybutyrate, 4-hydroxybutyryl-CoA,
crotonyl-CoA, 3-hydroxybutyryl-CoA, and acetoacetyl-CoA, formations of which
are catalyzed by succinyl-CoA reductase, succinate semialdehyde reductase,
4-hydroxybutyryl-CoA synthetase, 4-hydroxybutyryl-CoA dehydratase, crotonyl-
CoA hydratase, 3-hydroxybutyryl-CoA dehydrogenase, and acetoacetyl-CoA
β-ketothiolase, respectively [60].

12.4 Enzymes for CO2 Sequestration

Carbonic anhydrase (CA) is zinc metalloenzyme which catalyzes conversion of free


CO2 into bicarbonates and protons. It is one of the fastest known enzymes catalyz-
ing 104–106 reactions per second [73]. It has been found in a wide number of living
beings such as plants, animals, and microorganisms as they have been found
growing well in CO2-rich conditions, and CA was found essential for this purpose
12 Carbon Dioxide Sequestration by Biological Processes 327

[74, 75]. In human body, CA is present in the erythrocyte and converts poorly
soluble CO2 in aqueous solution, such as blood plasma to water-soluble bicarbonate
(HCO3) anion [76]. Human isozyme HCA II having molecular mass of 30,000 is
the fastest CA known so far, having a hydration rate of 1.4  106 molecules of CO2
per second per molecule of CA [19]. Some of the microalgae which have been
found containing CA are Chlamydomonas reinhardtii, Scenedesmus obliquus,
Dunaliella tertiolecta, Chlorella saccharophila, Chlorella vulgaris, Chlorella
pyrenoidosa, and Chlorococcum littorale [77]. BCA (bovine carbonic anhydrase)
has been found stable in wide range of pH (5–10) and temperature (up to 70  C)
[9]. Sulfate and zinc have been found to enhance CA activity.
There are contradictory reports on the effect of CO2 concentration on the activity
of CA in photosynthetic microorganisms such as plants, green algae, and
cyanobacteria. In most of the reports, activity of CA isolated from algae is inversely
proportional to the CO2 concentration. It was explained with the fact that RuBisCO
can catalyze both oxygenase and carboxylase activities depending upon the nearby
CO2 concentration [29]. CA helps in increasing the concentration of CO2 near the
RuBisCO site. CA can enhance the internal CO2 concentration up to 1,000 times
higher than the external fluid. Therefore, at lower CO2 concentration, expression of
CA increases to maintain carboxylase activity of RuBisCO which in turn reduces
atmospheric CO2 into cellular constituents such as starch, lipid, and protein.
However, in few reports, activity of CA initially increases with increase in CO2
concentration and later starts decreasing with further increase in CO2 concentration,
thus making bell-shaped curve [20]. The reason behind the decreasing trend of CA
at higher CO2 was postulated due to feedback inhibition by bicarbonate and/or
decrease in pH at this concentration.
Extracellular and intercellular crude extract of enzyme CA II isolated from
Chlorella vulgaris were 72 and 160 mg CaCO3 per milligram of protein, comparable
to 225 mg CaCO3 per milligram of protein with the purified enzyme from Citrobacter
freundii [78].

12.4.1 Mechanism of CO2 Captured by CA

Hydration of CO2 into solid carbonate such as calcium carbonate is a natural


process. There are five reactions required for the transformation of CO2 into
minerals [9, 19, 76].
1. Dissolution of gaseous CO2 in liquid (Eq. 12.12)
2. Hydration of aqueous CO2 into carbonic acids (H2CO3) (Eq. 12.13)
3. Dissociation of carbonic acid into bicarbonate ions (HCO32) and protons
(Eq. 12.16)
4. Dissociation of bicarbonate ions into carbonate ions (CO3) (Eq. 12.17)
5. Reaction of carbonate ions and calcium to form solid calcium carbonate
(Eq. 12.18)
328 K. Kumar and D. Das

12.4.1.1 CO2 Dissolution

CO2ðgÞ $ CO2ðaqÞ ð12:12Þ

12.4.1.2 Carbonic Acid Formation

CO2ðaqÞ þ H2 O $ H2 CO3 ð12:13Þ


k2 6:2  102 s1
K2 ¼ ¼ ¼ 2:6  103
k2 23:7 s1

Where k2, k2 and K2 are forward rate, reverse rate and equilibrium constants
respectively. As shown in Eq. (12.13), hydration of CO2 to carbonic acids is the
rate limiting step having very less forward reaction constant 6.2  102 s at 25  C
[18]. However, catalysis by CA increases the rate of reaction manifold showing the
high substrate specificity of this enzyme [19]. Enzyme catalyzes the CO2 hydration
reaction in the following two half reactions (Eqs. 12.14 and 12.15) [79]:

E  Zn  H2 O $ E  Zn þ OH þ Hþ ð12:14Þ

E  Zn  OH þ CO2 $ E  Zn  CO3  þ Hþ ! E  Zn  H2 O þ CO3  ð12:15Þ

The above reaction was found dominating when the pH is higher than 10 while it
was found negligible at pH less than 8 [80].

12.4.1.3 Bicarbonate Formation

H2 CO3 $ Hþ þ HCO3  ð12:16Þ


6 1
k3 8  10 s
K3 ¼ ¼ ¼ 1:7  104
k3 4:7  1010 s1

Where k3, k3 and K3 are forward rate, reverse rate and equilibrium constants respec-
tively. Bicarbonate formation reaction is diffusion controlled and very fast in nature.

12.4.1.4 Carbonate Formation

HCO3  $ Hþ þ CO3 2 ð12:17Þ


6 1
k4 8  10 s
K4 ¼ ¼ ¼ 4:69  1011
k4 4:7  1010 s1

Where k4, k4 and K4 are forward rate, reverse rate and equilibrium constants
respectively.
12 Carbon Dioxide Sequestration by Biological Processes 329

12.4.1.5 Calcium Carbonate Reaction

Ca2þ þ CaCO3 2 $ CaCO3 # ð12:18Þ

Calcium carbonate precipitates quickly at the saturation concentration of calcium


and carbonate ions. Therefore, continuous supply of carbonate ions by hydration of
CO2 and water is requisite [19]. CA was also found to enhance the calcium
carbonate precipitation reaction [9]. However, contrary to enhanced hydration
reaction, higher concentration of CA did not increase CaCO3 precipitation
reaction [9].
Some of the approaches which have been applied for improving the performance
of CA are (1) isolating CAs from thermophilic microorganisms, (2) use of protein
engineering to create thermotolerant enzymes, and (3) immobilizing the enzyme for
stabilization and confinement to cooler regions and process modification that
minimize the stresses such as cooling of the flue gas [81].

12.4.2 Immobilization of CA

Use of CA in aqueous solution has many disadvantages such as reusability and


recovery. Immobilization of the CA in the nanoparticles significantly improves
the catalytic property, storage, and thermal stability of the enzyme. Immobilized
BCA and HCA enzymes were found retaining nearly 90 % of enzymatic activity
for more than 20 cycles. Carbonic anhydrase (CA) can be immobilized on
functionalized and metal nanoparticles confined mesoporous silica for CO2
hydration and its sequestration to CaCO3 [76, 82]. Surface-modified magnetic
nanoparticle is one of such attractive templates for enzyme immobilization.
Enzyme can be easily recovered from a reaction medium by applying a static
magnetic field near the immobilized CA in the reactor [76, 82]. Inert materials
such as chitosan and sodium alginate are widely used for immobilization of
enzymes and microorganism. Whole cells of Pseudomonas fragi, Micrococcus
lylae, and Micrococcus luteus 2 were immobilized on different biopolymer
matrices [83]. Bovine carbonic anhydrase (BCA) was covalently immobilized
by Vinoba et al. (2012) onto OAPS (octa(aminophenyl)silsesquioxane)-
functionalized Fe3O4/SiO2 nanoparticles by using glutaraldehyde as a spacer
[76]. Immobilization of CA can be done in many ways such as adsorption on
surfaces, entrapment within matrices, and cross-linking within polymeric scaffold
[80]. Polyurethane foam is a highly porous hydrophilic polymeric material where
enzyme immobilization was easy and fast. CA had 100 % activity over time, and
thus reusability was not a concern [80].
330 K. Kumar and D. Das

12.4.3 Application of CA

Bovine CA has been proposed to inject into wellbore of geological formations to


prevent CO2 leakage through it [84]. CO2 sequestration by mineral carbonation can
use CA to make the process feasible at large scale. Alkaline silicates are abundant
and high enough to sequester all the CO2 emitted from total fossil fuels. Alkaline
silicates can dissolve to provide cations in acidic conditions. However, high
alkalinity is required for the increase of the rate of gaseous CO2 dissolving into
the carbonate ions CO32. Therefore, enhancing the release of divalent cations
from the alkaline silicates and enhancing alkalinity are some of the challenges of
the process. Increase in alkalinity increases the rate of dissolution of CO2 into
the carbonate ions CO32, and acidic conditions release the divalent cations
required for the formation of carbonates [17]. Denitrification, methane production,
and sulfate reduction are some of the alkalinity-producing metabolic processes. There-
fore, these processes enhance the carbonation process. Integrating acid-producing
process for silicate dissolution and alkaline-producing process for carbonate
precipitation together can be used for CO2 mitigation using biological mineral
carbonation. Dupraz et al. (2009) has experimentally shown the use of Bacillus
pasteurii as a model carbonate precipitating bacteria on the geological sequestration
of CO2 and its transformation into solid carbonate phases [85].

12.4.4 Challenges in Use of CA for CO2 Sequestration

Enzyme works better at optimum temperature, pH, and enzyme concentration


[83]. At lower pH, carbonates prefer to be in dissolved state than the precipitated
form, while at higher pH, carbonate ions form but concerns about economical and
environmental aspect arise [19]. Lifetime and activity of enzyme greatly depend upon
pH, temperature, other ions such as CN, and higher concentration of SOx and NOx
[19]. Higher cost of enzyme and its large-scale production are other bottlenecks to
overcome to make the process successful in reality.

12.5 Conclusions

Imbalance between CO2 emission and sink is the reason for steep rise in the earth’s
atmospheric CO2 concentration. This may be the reason for the rise in global mean
temperature, causing melting of glaciers, rise in sea level, ocean acidification,
unpredicted climate changes, etc. Several biological processes are available for
CO2 mitigation. Carbonic anhydrase (CA) enzyme, algae, cyanobacteria, bacteria,
and terrestrial plants are some of the biological methods used for CO2 sequestration.
However, these processes are slow and limited in application. Identifying limiting
12 Carbon Dioxide Sequestration by Biological Processes 331

parameters and application of technology can accelerate the natural processes man-
ifold. For example, CO2 hydration, the limiting step in the transformation of gaseous
CO2 to solid bicarbonates, can be accelerated by catalyzing the reaction using
CA. Forest and oceanic fertilization can be applied to enhance the photosynthetic
efficiency of terrestrial plants and oceanic phytoplanktons, respectively. Similarly,
algae and cyanobacteria can be exploited for CO2 sequestration as they can be grown
efficiently at higher CO2 concentration with higher photosynthetic efficiency.

Acknowledgments The authors gratefully acknowledge the Council of Scientific and Industrial
Research (CSIR), Govt. of India, for senior research fellowship and Department of Biotechnology
(DBT) and Ministry of New and Renewable Energy (MNRE), Govt. of India, for the financial
support. The authors also acknowledge Ms. Pallavi Sinha for reviewing the manuscript.

References

1. Tans P (2010) Trends in atmospheric carbon dioxide. In: Earth System Research Laboratory
Global Monitoring Division. http://www.esrl.noaa.gov/gmd/ccgg/trends. Accessed 15 May 2013
2. http://climate.nasa.gov/evidence/
3. Bardgett RD, Freeman C, Ostle NJ (2008) Microbial contributions to climate change through
carbon cycle feedbacks. ISME J 2:805–814
4. Stewart C, Hessami M (2005) A study of methods of carbon dioxide capture and
sequestration–the sustainability of a photosynthetic bioreactor approach. Energy Convers
Manage 46:403–420
5. Sharma A, Bhattacharya A, Shrivastava A (2011) Biomimetic CO2 sequestration using
purified carbonic anhydrase from indigenous bacterial strains immobilized on biopolymeric
materials. Enzyme Microb Technol 48:416–426
6. Oelkers EH, Cole DR (2008) Carbon dioxide sequestration: a solution to a global problem.
Elements 4:305–310
7. Kumar K, Dasgupta CN, Nayak B et al (2011) Development of suitable photobioreactors for
CO2 sequestration addressing global warming using green algae and cyanobacteria. Bioresour
Technol 102:4945–4953
8. Lee JW, Li R (2001) A novel strategy for CO2 sequestration and clean air protection. In: Pro-
ceedings of fire national conference on carbon sequestration, Washington, DC, 14–17 May 2001
9. Mirjafari P, Asghari K, Mahinpey N (2007) Investigating the application of enzyme carbonic
anhydrase for CO2 sequestration purposes. Ind Eng Chem Res 46:921–926
10. Bachu S (2000) Sequestration of CO2 in geological media: criteria and approach for site
selection in response to climate change. Energy Convers Manage 41:953–970
11. Falkowski P, Scholes RJ, Boyle E et al (2000) The global carbon cycle: a test of our knowledge
of earth as a system. Science 290:291–296
12. Hall DO, Woods J, House J (1992) Biological systems for uptake of carbon dioxide. Energy
Convers Manage 33(5–8):721–728
13. Whitmarsh J, Govindjee (1999) The photosynthetic process. In: Singhal GS, Renger G, Sopory
SK, Irrgang K-D, Govindjee (eds) Concepts in photobiology: photosynthesis and photomorpho-
genesis. Narosa Publishers/Kluwer, New Delhi/Dordrecht, pp 11–51
14. Benemann JR (1993) Utilization of carbon dioxide from fossil fuel-burning power plants with
biological systems. Energy Convers Manage 34(9–11):999–1004
15. Skjånes K, Lindblad P, Muller J (2007) BioCO2 – a multidisciplinary, biological approach
using solar energy to capture CO2 while producing H2 and high value products. Biomol Eng
24:405–413
332 K. Kumar and D. Das

16. Loubiere K, Olivo E, Bougaran G et al (2009) A new photobioreactor for continuous


microalgal production in hatcheries based on external-loop airlift and swirling flow.
Biotechnol Bioeng 102(1):132–147
17. Salek SS, Kleerebezem R, Jonkers HM et al (2013) Mineral CO2 sequestration by environmental
biotechnological processes. Trends Biotechnol 31(3):139–146
18. Druckenmiller ML, Maroto-Valer MM (2005) Carbon sequestration using brines of adjusted
pH to form mineral carbonates. Fuel Process Technol 86:1599–1614
19. Bond GM, Stringer J, Brandvold DK et al (2001) Development of integrated system for
biomimetic CO2 sequestration using the enzyme carbonic anhydrase. Energy Fuels
15:309–316
20. Ramanan R, Kannan K, Sivanesan SD et al (2009) Bio-sequestration of carbon dioxide using
carbonic anhydrase enzyme purified from Citrobacter freundii. World J Microbiol Biotechnol
25:981–987
21. Beedlow PA, Tingey DT, Phillips DL et al (2004) Rising atmospheric CO2 and carbon
sequestration in forests. Front Ecol Environ 2(6):315–322
22. Norby RJ, DeLucia EH, Gielen B et al (2005) Forest response to elevated CO2 is conserved
across a broad range of productivity. Proc Natl Acad Sci U S A 102:18052–18056
23. Norby RJ, Wullschleger SD, Gunderson CA (1999) Tree responses to rising CO2 in field
experiments: implications for the future forest. Plant Cell Environ 22:683–714
24. Griffin KL, Seemann JR (1996) Plants, CO2 and photosynthesis in the 21st century. Chem Biol
3(4):245–254
25. Karnosky DF (2003) Impacts of elevated atmospheric CO2 on forest trees and forest ecosystems:
knowledge gaps. Environ Int 29:161–169
26. Tognetti R, Cherubini P, Innes JL (2000) Comparative stem growth rates of Mediterranean
trees under background and naturally enhanced ambient CO2 concentrations. New Phytol
146:59–74
27. Idso SB, Idso KE, Garcia RL et al (1995) Effects of atmospheric CO2 enrichment and foliar
methanol application on net photosynthesis of sour orange tree (Citrus aurantium; Rutaceae)
leaves. Am J Bot 82:26–30
28. Manning DAC, Renforth P (2013) Passive sequestration of atmospheric CO2 through coupled
plant-mineral reactions in urban soils. Environ Sci Technol 47:135–141
29. Moroney JV, Somanchi A (1999) How do algae concentrate CO2 to increase the efficiency of
photosynthetic carbon fixation. Plant Physiol 119:9–16
30. Tsai DDW, Ramaraj R, Chen PH (2012) Growth condition study of algae functions in
ecosystem for CO2 bio-fixation. J Photochem Photobiol B Biol 107:27–34
31. Yang H, Xu Z, Fan M et al (2008) Progress in carbon dioxide separation and capture: a review.
J Environ Sci 20:14–27
32. Oren R, Ellsworth DS, Johnsen KH et al (2001) Soil fertility limits carbon sequestration by
forest ecosystems in a CO2-enriched atmosphere. Nature 411:469–471
33. Yuan H, Ge T, Wu X et al (2012) Long-term field fertilization alters the diversity of
autotrophic bacteria based on the ribulose-1,5-biphosphate carboxylase/oxygenase (RuBisCO)
large-subunit genes in paddy soil. Appl Microbiol Biotechnol 95:1061–1071
34. Munoz R, Guieysse B (2006) Algal-bacterial processes for the treatment of hazardous
contaminants: a review. Water Res 40:2799–2815
35. Kumar K, Sirasale A, Das D (2013) Use of image analysis tool for the development of light
distribution pattern inside the photobioreactor for the algal cultivation. Bioresour Technol
143:88–95
36. Hanagata N, Takeuchi T, Fukuju Y et al (1992) Tolerance of microalgae to high CO2 and high
temperature. Phytochemistry 31:3345–3348
37. Carvalho AP, Meireles LA, Malcata FX (2006) Microalgal reactors: a review of enclosed
systems design and performances. Biotechnol Prog 22:1490–1506
38. Lee JS, Lee JP (2003) Review of advances in biological CO2 mitigation technology.
Biotechnol Bioprocess Eng 8:354–359
12 Carbon Dioxide Sequestration by Biological Processes 333

39. Castenholz RW, Waterbury JB (1989) Group I. Cyanobacteria. In: Staley JT, Bryant MP,
Pfennig N, Holt JG (eds) Bergey’s manual of systematic bacteriology, vol 3. Williams &
Wilkins, Baltimore, pp 1710–1727
40. Garofalo R (2009) Algae and aquatic biomass for a sustainable production of 2nd generation
biofuels. AquaFUELs Taxon Biol Biotechnol, pp 1–258
41. Garlick S, Oren A, Padan E (1977) Occurrence of facultative anoxygenic photosynthess
among unicellular and filamentous cyanobacteria. J Bacteriol 129:623–629
42. Srirangan K, Pyne ME, Chou CP (2011) Biochemical and genetic engineering strategies to
enhance hydrogen production in photosynthetic algae and cyanobacteria. Bioresour Technol
102:8589–8604
43. Giordano M, Beardall J, Raven JA (2005) Mechanisms in algae: mechanisms, environmental
modulation, and evolution. Annu Rev Plant Biol 56:99–131
44. Zeng X, Danquah MK, Chen XD et al (2011) Microalgae bioengineering: from CO2 fixation to
biofuel production. Renew Sust Energy Rev 15:3252–3260
45. Cannon GC, Heinhorst S, Kerfeld CA (2010) Carboxysomal carbonic anhydrases: structure
and role in microbial CO2 fixation. Biochim Biophys Acta 1804:382–392
46. Coleman JR (1991) The molecular and biochemical analysis of CO2 concentrating mechanisms
in cyanobacteria and microalgae. Plant Cell Environ 14:861–867
47. Kaplan A, Reinhold L (1999) CO2 concentrating mechanisms in photosynthetic microorganisms.
Annu Rev Plant Physiol Plant Mol Biol 50:539–570
48. Leegood RC (2002) C4 photosynthesis: principles of CO2 concentration and prospects for its
introduction into C3 plants. J Exp Bot 53(369):581–590
49. Badger MR, Price GD (2003) CO2 concentrating mechanisms in cyanobacteria: molecular
components, their diversity and evolution. J Exp Bot 54(383):609–622
50. Price GD, Maeda S, Omata T et al (2002) Modes of active inorganic carbon uptake in the
cyanobacterium Synechococcus sp. PCC7942. Funct Plant Biol 29:131–149
51. Spalding MH (2008) Microalgal carbon dioxide concentrating mechanisms: Chlamydomonas
inorganic carbon transporters. J Exp Biol 59(7):1463–1473
52. Hu JJ, Wang L, Zhang SP et al (2011) Enhanced CO2 fixation by a non-photosynthetic
microbial community under anaerobic conditions: optimization of electron donors. Bioresour
Technol 102:3220–3226
53. Nelson DL, Cox MM (2005) Lehninger principles of biochemistry, 4th edn. W.H. Freeman &
Co, New Jersey
54. Madigan MT, Martinko JM, Parker J (2003) Brock: biology of microorganisms. Prentice Hall,
Upper Saddle River
55. Calvin M, Benson AA (1948) The path of carbon in photosynthesis. Science 107:476–480
56. Kumar K, Das D (2012) Growth characteristics of Chlorella sorokiniana in airlift and bubble
column photobioreactors. Bioresour Technol 116:307–313
57. Kumar K, Roy S, Das D (2013) Continuous mode of carbon dioxide sequestration by
C. sorokiniana and subsequent use of its biomass for hydrogen production by E. cloacae
IIT-BT 08. Bioresour Technol 145:116–122
58. Fast AG, Papoutsakis ET (2012) Stoichiometric and energetic analyses of non-photosynthetic
CO2-fixation pathways to support synthetic biology strategies for production of fuels and
chemicals. Curr Opin Chem Eng 1:1–16
59. Müller V (2003) Bacteria energy conservation in acetogenic bacteria. Appl Environ Microbiol
69(11):6345–6353
60. Saini R, Kapoor R, Kumar R et al (2011) CO2 utilizing microbes – a comprehensive review.
Biotechnol Adv 29:949–960
61. Atomi H (2002) Microbial enzymes involved in carbon dioxide fixation. J Biosci Bioeng
94(6):497–505
62. Evans MC, Buchanan BB, Arnon DI (1966) A new ferredoxin-dependent carbon reduction
cycle in a photosynthetic bacterium. Proc Natl Acad Sci U S A 55:928–934
63. Sintsov NV, Ivanovskii RN, Kondrateva EN (1980) ATP-dependent citrate lyase in the green
phototrophic bacterium, Chlorobium limicola. Mikrobiologiia 49:514–516
334 K. Kumar and D. Das

64. Shiba H, Kawasumi T, Igarashi Y et al (1985) The CO2 assimilation via the reductive
tricarboxylic acid cycle in an obligately autotrophic, aerobic hydrogen-oxidizing bacterium,
Hydrogenobacter thermophilus. Arch Microbiol 141:198–203
65. Alber BE, Fuchs G (2002) Propionyl-coenzyme A synthase from Chloroflexus aurantiacus, a
key enzyme of the 3-hydroxypropionate cycle for autotrophic CO2 fixation. J Biol Chem
277:12137–12143
66. Burton NP, Williams TD, Norris PR (1999) Carboxylase genes of Sulfolobus metallicus. Arch
Microbiol 172:349–353
67. Strauss G, Fuchs G (1993) Enzymes of a novel autotrophic CO2 fixation pathway in the
phototrophic bacterium Chloroflexus aurantiacus, the 3-hydroxypropionate cycle. Eur J Biochem
215:633–643
68. Herter S, Fuchs G, Bacher A et al (2002) A bicyclic autotrophic CO2 fixation pathway in
Chloroflexus aurantiacus. J Biol Chem 277:20277–20283
69. Hügler M, Menendez C, Schagger H et al (2002) Malonyl-coenzyme A reductase from
Chloroflexus aurantiacus, a key enzyme of the 3-hydroxypropionate cycle for autotrophic
CO2 fixation. J Bacteriol 184:2404–2410
70. Berg IA, Ramos-Vera WH, Petri A et al (2010) Study of the distribution of autotrophic CO2
fixation cycles in Crenarchaeota. Microbiology 156:256–259
71. Huber H, Gallenberger M, Jahn U et al (2008) A dicarboxylate/4-hydroxybutyrate autotrophic
carbon assimilation cycle in the hyperthermophilic archaeum Ignicoccus hospitalis. Proc Natl
Acad Sci U S A 105:7851–7856
72. Ramos-Vera WH, Berg IA, Fuchs G (2009) Autotrophic carbon dioxide assimilation in
Thermoproteales revisited. J Bacteriol 191:4286–4297
73. Heck RW, Tanhauser SM, Manda R et al (1994) Catalytic properties of mouse carbonic
anhydrase V. J Biol Chem 269:24742–24746
74. Karlsson J, Clarke AK, Chen ZY et al (1998) A novel a-type carbonic anhydrase associated
with the thylakoid membrane in Chlamydomonas reinhardtii is required for growth at ambient
CO2. EMBO J 17(5):1208–1216
75. Kusian B, Sultemeyer D, Bowien B (2002) Carbonic anhydrase is essential for growth of
Ralstonia eutropha at ambient CO2 concentrations. J Bacteriol 184(18):5018–5026
76. Vinoba M, Bhagiyalakshmi M, Jeong SW et al (2012) Carbonic anhydrase immobilized on
encapsulated magnetic nanoparticles for CO2 sequestration. Chem Eur J 18:12028–12034
77. Miyachi S, Iwasaki I, Shiraiwa Y (2003) Historical perspective on microalgal and
cyanobacterial acclimation to low and extremely high CO conditions. Photosynth Res
77:139–153
78. Li L, Fu M, Zhao Y et al (2012) Characterization of carbonic anhydrase II from Chlorella
vulgaris in bio-CO2 capture. Environ Sci Pollut Res 19:4227–4232
79. Brauer M, Perez-Lustres JL, Weston J (2002) Quantitative reactivity model for the hydration
of carbon dioxide by biomimetic zinc complexes. Inorg Chem 41:1454–1463
80. Ozdemir E (2009) Biomimetic CO2 sequestration: 1. Immobilization of carbonic anhydrase
within polyurethane foam. Energy Fuels 23:5725–5730
81. Savile CK, Lalonde JJ (2011) Biotechnology for the acceleration of carbon dioxide capture and
sequestration. Curr Opin Biotechnol 22:818–823
82. Vinoba M, Yoon Y, Nam SC et al (2012) CO2 conversion to CaCO3 by immobilized carbonic
anhydrase. KIC News 15(2):32–40
83. Sharma N, Nayak RK, Dadhwal VK et al (2011) Diurnal and seasonal variation of measured
atmospheric CO2 at Dehradun during 2009. In: International archives of the photogrammetry,
remote sensing and spatial information sciences, volume XXXVIII-8/W20, 2011 ISPRS
Bhopal 2011 workshop, Bhopal, 8 Nov 2011
84. Mahinpeya N, Asghari K, Mirjafari P (2011) Biological sequestration of carbon dioxide in
geological formations. Chem Eng Res Des 89:1873–1878
85. Dupraz S, Ménez B, Gouze P et al (2009) Experimental approach of CO2 biomineralization in
deep saline aquifers. Chem Geol 265:54–62
Part IV
Reactions in/Under Dense Phase
Carbon Dioxide
Chapter 13
Homogeneous Catalysis Promoted
by Carbon Dioxide

Ran Ma, Zhen-Feng Diao, Zhen-Zhen Yang, and Liang-Nian He

13.1 Introduction

The global increase in environmental awareness has led to an ever-increasing


control over the use and disposal of hazardous materials by the chemical industry
and searching for “cleaner” alternatives for organic synthetic processes. One
obvious target area is the replacement of conventional organic solvents by suitable
candidates, such as supercritical fluids (particularly CO2), ionic liquids [1],
water [2], or even solvent-free systems [3]. Supercritical CO2 (scCO2), which is
recognized as CO2 heated and pressurized beyond its critical point (Tc ¼ 31.06  C,
Pc ¼ 7.38 MPa), is considered to offer numerous advantages as reaction medium
offering safe environment owing to its unique physical properties, such as abun-
dantly available and cheap; nontoxic and environmentally benign even under
oxidative conditions; high gaseous miscibility with hydrogen, oxygen, carbon
monoxide, and methane; effective mass transfer; easily tunable properties with
variation of pressure or temperature; and weakening of the solvent interactions
around the reacting species [4–7]. In addition, surface tension effects can be
eliminated since scCO2 has negligible surface tension like a gas [8, 9]. Most
importantly, separation of products and solvent (scCO2) can easily be performed
by depressurization, and the separated CO2 can be reused by pressurization. On the
other hand, in the case of aerobic oxidation reactions, the miscibility of organic
substrates and catalysts with O2 is crucial to achieve high oxidation rate, but the
metal component of catalysts tends to undergo overoxidation, thus leading to the
catalyst deactivation [10]. Hence, dense-phase (liquid or supercritical) CO2 has
been recently recognized as an innovative and environmentally friendly reaction
medium for chemical syntheses and especially for the metal-catalyzed processes,
mainly as an alternative to conventional organic solvents [11, 12]. This chapter

R. Ma • Z.-F. Diao • Z.-Z. Yang • L.-N. He (*)


State Key Laboratory of Elemento-Organic Chemistry, Nankai University, Tianjin 300071,
People’s Republic of China
e-mail: heln@nankai.edu.cn

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 337
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_13,
© Springer-Verlag Berlin Heidelberg 2014
338 R. Ma et al.

summarizes the utilization of scCO2 as a solvent and promoter in organic reactions


including hydrogenation, carbonylation, C–C forming reaction, oxidation reaction,
and polymerization.

13.2 Hydrogenation

13.2.1 Asymmetric Hydrogenation

The monodentate phosphites containing ortho- and meta-closo-dodecarboranyl


groups have widely been used as chiral ligands in the Rh-catalyzed asymmetric
hydrogenation of dimethyl itaconate with high enantioselectivities (up to 93 % ee in
scCO2) [13, 14]. In order to achieve the comparable enantioselective with bidentate
phosphines (97–99 % ee), the easily accessible monodentate phosphite-type
ligands, for instance, PipPhos (L1) and MorfPhos (L2), are successfully applied
to the hydrogenation of dimethyl itaconate 1 and methyl 2-acetamidoacrylate 3 in
scCO2 with high enantioselectivity (up to 99 % ee) and complete conversion within
35–50 min as depicted in Scheme 13.1 [15]. The high reaction rates could

O O O
P N P N O P O
O O O

L1 L2 L3
O O
O [Rh(COD)2BF4/2Ln] O
*
O H2, scCO2 O
O O
1 2
L1: 97.8% ee of R
L2: 98.3% ee of R
L3: 92.0% ee of R
O O

O [Rh(COD)2BF4/2Ln] O
*
HN O H2, scCO2 HN O

O O
3 4
L1: 99.0% ee of S
L2: 97.0% ee of S

Scheme 13.1 Rhodium-catalyzed hydrogenation of prochiral olefins in scCO2 [15]


13 Homogeneous Catalysis Promoted by Carbon Dioxide 339

presumably be attributed to the higher miscibility and diffusivity of gaseous


hydrogen in the supercritical medium, thereby eliminating mass transfer resistance
compared to that in CH2Cl2 [15, 16].
A successive application of monodentate phosphoramidite ligands has been
covered successfully in the Rh-catalyzed hydrogenation of α-dehydroamino acid
derivatives in scCO2 with high enantioselectivity (up to 97 % ee) and excellent
reaction rates [16]. Recently, sterically congested carborane-containing mono-
dentate ligands (L3, Scheme 13.1) have been designed and synthesized as chiral
ligands in Rh-catalyzed hydrogenation of prochiral olefins in scCO2 [17]. High
enantioselectivity (up to 92 % ee) and reactivities (100 % conversion in 45–60 min)
are obtained when hydrogen pressure and total pressure are adjusted within H2
8 MPa and total pressure 20 MPa [17].
The Ru complex is also found to be an efficient catalyst for the hydrogenation
reactions in the environmentally friendly solvents [18–20]. While the solubility
issues for most aromatic substrates, reagents, ligands, and catalysts in scCO2
represented a critical factor for this type of asymmetric hydrogenation reaction, in
this context, cosolvent, such as alcohol, is necessary for some substrates in order to
increase the solubility in scCO2 [21]. For example, the fluorinated alcohols such
as CF3(CF2)6CH2OH can increase both enantioselectivity (up to 89 % ee) and
reactivity (over 99 % yield) for the hydrogenation of tiglic acid in scCO2 to
afford 2-methylbutanoic acid catalyzed by (S)-H8-BINAP-Ru(II) [22]. On the
other hand, perfluoroalkylBINAP is successfully used as a soluble chiral ligand
for the asymmetric hydrogenation in scCO2 with trifluorotoluene as cosolvent
(Scheme 13.2) [23].

H NHAc H NHAc
Ru-catalyst (5a or 5b)
H * H
H3C CO2Me H2 (3 MPa), CO2 (10 MPa)
H3C CO2Me
trifluorotoluene 0.5 mL
50 °C, 5 h
Conversion 100%
ee 74%

Rf Rf

P P

P P

Rf Rf

5a Rf= C6F13 or C10F21 5b Rf= C6F13 or C10F21

Scheme 13.2 The perfluoroalkylBINAP ligands for the asymmetric hydrogenation [23]
340 R. Ma et al.

R R
N HN
Ir-cat* 6, H2 (3 MPa)
3 Ph (1)
Ph CH3 scCO2(d=0.75 g/mL ) CH3
40 °C, 20 h
ee 80%

N Ir(COD2)BARF)/ 2 L1 HN
H2 (4 MPa), scCO2 (2)
45 °C, 2 h *
ee 95%

C6F13(CH2)2
CF3

N
P
B
Ir
4
CF3
C6F13(CH2)2

Scheme 13.3 Ir-catalyzed hydrogenation of acyclic arylimines in scCO2 [25, 26]

The enantioselective hydrogenation of imines provides an attractive route to


the preparation of optical-active amines and has attracted much attention in both
academic and industrial research [24]. Leitner has showed that the cationic iridium
complex 6 with chiral phosphinodihydrooxazoles, modified with perfluoroalkyl
groups for enhancing the solubility in scCO2, has been tested in the hydrogena-
tion of N-(1-phenylethylidene)aniline in scCO2, resulting in ee of up to 80 %
(Scheme 13.3 (1)) [25]. Recently, the monodentate phosphite-type ligands have also
been successfully used in the Ir-catalyzed asymmetric hydrogenations of acyclic
arylimines in scCO2 (Scheme 13.3 (2)) [26]. High yields (100 % conversion in
50–120 min) and enantioselectivities (up to 95 %) can be obtained under the
optimal conditions (4 MPa H2, 25 MPa total pressure, 45  C) when L1
(Scheme 13.1) is used as chiral ligands. It is worth noting that electron-donating
or electron-withdrawing groups at the para-position of the N-aryl ring almost have
no effect on the enantioselectivity [26].
ScCO2 has also been demonstrated as a benign solvent for the asymmetric
homogeneous hydrogenation of carbon–oxygen double bond [27]. A ruthenium
complex, i.e., [RuCl2(C6H6)]2-(R)-BINAP derived in situ from the ruthenium
source [RuCl2(C6H6)]2 and (R)-BINAP, is successfully found to be active for the
13 Homogeneous Catalysis Promoted by Carbon Dioxide 341

O O OH O
[RuCl2(C6H6)]2-(R)-BINAP
Cl Cl
OEt H2 2.7MPa, CO2-[bmim]BF4 OEt
7 75 °C, 2 h 8
ee 97%

MeOOC MeOOC
O
COOMe [RuCl2(C6H6)]2-(R)-BINAP
O + O
O O
H2, CO2-[bmim]BF4
COOMe 75 °C, 13 h (2R, 3S) (2R, 3R)
10a 10b
9
ee 90.5% ee 83%

Scheme 13.4 Asymmetric hydrogenation of keto esters by Ru–BINAP catalyst [28]

11 or 12 or 13
H2, scCO2

F3C F3C

P Rh/Cl P Rh/BArF P Rh/BArF

F3C 3 3 F3C 3 3 3 3

11 12 13

Scheme 13.5 Hydrogenation of styrene catalyzed by fluorinated rhodium–phosphine complexes [32]

asymmetric hydrogenation of ethyl 4-chloro-3-oxobutyrate 7 and dimethyl


acetylsuccinate 9 with high enantioselectivity [28]. Interestingly, when the ionic
liquid [bmim]BF4 is added as a polar cosolvent, the hydrogenation enantios-
electivity increases accordingly to give the target hydroxy ester 8 and lactones 10
(Scheme 13.4).

13.2.2 Hydrogenation of Styrene

Fluorine groups attached to ligands and fluorinated solvents such as scCH2F2


can increase the solubility of the catalyst containing fluorinated substituents in
scCO2 [29]. BArF counterion and CF3 groups are commonly used in asymmetric
hydrogenation reactions with an increasing effect on solubility in the scCO2 [25, 30,
31]. Three different fluorinated rhodium–phosphine complexes, [P(Ph(CF3)2)3]3RhCl
11, [P(Ph(CF3)2)3]3RhBArF 12, and [P(Ph)3]3RhBArF 13, which are observed to
be soluble in scCO2, thus are developed as efficient homogeneous catalysts for
the hydrogenation of styrene (Scheme 13.5) [32]. The synthesized Rh-catalyst 12
342 R. Ma et al.

Scheme 13.6 Ru-catalyzed Bcat


hydroboration of styrene
derivatives to boronate HBcat, Rh/L (1:2)
esters in scCO2 [35] scCO2
MeO MeO
40 °C, 5 h
14 Yield 100%

is the most effective catalyst with a conversion of 77.4 %, owing to the


electron-withdrawing effect of fluoroalkyl moieties which presumably could have
an important impact on the catalytical efficiency.

13.2.3 Hydroboration of Styrene

Boronate esters are one of the most extensively used intermediates in organic
synthesis and pharmaceutical chemistry [33, 34]. Tumas and co-workers have
demonstrated the first production of boronate esters from hydroboration of styrene
derivatives in scCO2 with a rhodium precursor and phosphorus ligands [35].
Incorporation of fluorinated functional groups into phosphine-based ligands can
not only boost the solubility of the corresponding metal complexes but also control
the stereoelectronic environment of catalysts [36]. Alkylboronate ester 14 is found
as the unique product when the cyclohexyl-substituted phosphine PR2(ORF) (L/Rh
ratio ¼ 2:1) is employed as a ligand, while a mixture of alkylboronate ester isomers
are obtained when tetrahydrofuran (THF) is used as solvent instead of scCO2 under
the identical conditions (Scheme 13.6).

13.3 Carbonylation

13.3.1 Hydroformylation

Hydroformylation, which is adding one CO and H2 to a carbon–carbon double bond


to produce linear and branched aldehydes containing one more carbon number than
the starting olefin, is an important transformation process of significant interest for
academia as well as industry. Using scCO2 as a reaction medium for homogeneous
hydroformylation and as a means to tune selectivity or activity has received much
attention with the potential benefits as applicable to this system as they have been
shown to be with hydrogenation.
Rathke et al. have reported the first example of homogeneous hydroformylation
in scCO2 with a cobalt carbonyl catalyst (soluble in CO2 without the need of any
modification) by virtue of high-pressure NMR spectroscopy (Scheme 13.7)
[37]. Improved yield of linear to branched butyraldehyde products up to 88 % is
13 Homogeneous Catalysis Promoted by Carbon Dioxide 343

Scheme 13.7 HCo(CO)4


Hydroformylation of
propylene in scCO2 with H2
Co2(CO)8 [37]
cat.Co2(CO)8
CHO + CHO
CO, H2
scCO2 (d=0.5 g/mL3) Yield 88%
80 °C

Scheme 13.8 MnH(CO)5- Ph Ph Ph Ph


CO, H2
catalyzed hydroformylation + MnH(CO)5
H CHO
of cyclopropene [39] scCO2 (20 MPa)
60 °C, 3.5 h

Ph2P CO2C16H33
O O C9H19
P P C16H33O2C
3 3

Scheme 13.9 Phosphite ligand for hydroformylation of long chain alkenes [41]

obtained when the hydroformylation proceeds in CO2 at 80  C without stirring. The


rates of formation of both cobalt intermediate, i.e., HCo(CO)4, and aldehydes are
found to be comparable to reactions being performed in conventional nonpolar
solvents. Activation energy (23.3  1.4 kcal/mol) in scCO2 is comparable to those
measured in conventional organic solvents (27–35 kcal/mol) [38].
Subsequently, hydroformylation of 3,3-dimethyl-1,2-diphenylcyclopropene cat-
alyzed by MnH(CO)5 in scCO2 at 60  C has been developed by Jessop’s group, to
produce 66 % of alkane and 34 % of aldehydes, with the same chemoselectivity in
hexane and pentane (Scheme 13.8). The aldehydes are primarily formed through
nonradical pathways [39].
Simple trialkylphosphines, e.g., PEt3, which are readily available, give high
catalytic activity for the hydroformylation of hex-1-ene in scCO2 [40]. A slight
higher normal-to-branched ratio of aldehydes can be obtained taking [Rh2(OAc)4]
and PEt3 as catalyst precursors, while a similar rate is obtained in toluene. Further-
more, Sellin and Cole–Hamilton have developed an alternative approach to the
hydroformylation of long chain alkenes using insoluble phosphite-modified rho-
dium catalysts in scCO2 (Scheme 13.9) [41]. Rhodium catalyst prepared in situ
from Rh2(OAc)4 and P(OPh)3 gave reasonable activity and high regioselectivity.
The catalyst can be reused several times because the reaction products can be easily
removed from the reaction by flushing the reactor with scCO2.
Leitner et al. [29] as well as Erkey et al. [42] have independently investigated
hydroformylation of an olefin in scCO2 using fluorinated catalysts in order to
enhance the CO2 philicity and catalytic activity in scCO2. A CO2-soluble Rh
complex 15 with a polyfluoroalkyl-substituted triarylphosphine ligand is developed
344 R. Ma et al.

Rh, CO, H2 CHO


CHO
n-C6H13 + n-C6H13
scCO2 n-C6H13
CF3 CF3
Rh cat:
CO

F3C
O F3C P Rh P CF3
Rh(g4-C8H12) + P
O 3
F3C Cl
15 (CH2)2(CH2)5CF3
CF3 16 CF3

Scheme 13.10 Fluorinated catalysts for hydroformylation [29, 42]

Cat. 17 CHO + CHO


C5H11 CO, H2 C5H11 C5H11
scCO2

Cat 17:
F F
Co Co
Co

P Co Co P

F Co Co Co F

F F

Scheme 13.11 Hydroformylation of 1-octene in scCO2 by complex 17 [43]

for this purpose [29]. Hydroformylation of 1-octene gives a linear aldehyde in a


good yield with 82 % selectivity (Scheme 13.10). Side reactions such as hydroge-
nation or isomerization of the olefin can be effectively suppressed in scCO2. The
trans-RhCl(CO)(P(p-CF3C6H4)3)2 16 (Scheme 13.10) with moderate solubility in
scCO2 also shows high activity for the hydroformylation of 1-octene. In addition,
hydrogenation and isomerization products are insignificant, and the normal-to-
branched ratio of 2.4 is acceptable with the complete conversion of 1-octene.
A bi-cobalt complex, i.e., Co2(CO)6[P(3-FC6H4)3]2 17, is designed and prepared
for the hydroformylation of 1-octene in scCO2 as shown in Scheme 13.11 [43].
Substantial improvement of the selectivity toward aldehyde is reached by using 17
in comparison with Co2(CO)8. An additional excess of phosphine is unnecessary for
the conversion. Complex 17 can be efficiently separated from the reaction mixture
because it is insoluble in the cold reaction mixture while completely soluble in the
supercritical reaction medium.
Phosphite ligands to form rhodium catalysts give higher reactivity than phos-
phines [44]. Masdeu-Bultó and co-workers show soluble rhodium–phosphite cata-
lytic systems for hydroformylation of 1-octene in scCO2 [45]. A series of aryl
phosphite ligands (18–21) are prepared with different donor and steric properties,
13 Homogeneous Catalysis Promoted by Carbon Dioxide 345

Scheme 13.12 Structures


of phosphite ligands P O CF3 P O OCF3
containing CF3 group [45] 3 3
18 19

P O
P O
3
3
CF3
OCF3
20 21

Scheme 13.13 Structures


of phosphorus ligands P O P O
without perfluoroalkyl P
3 3 3
substituents [46]

22 23 24

and the introduction of a CF3 group could increase the π-acceptor ability of the
phosphite (Scheme 13.12). High-pressure IR and NMR spectroscopies are used to
investigate reactivity of rhodium precursors with these ligands in the presence of
CO and H2, showing that [RhH(CO)(18–21)3] is formed as the main species in
solution. High conversion (up to 93–99 %) for [Rh(acac)(CO)2]/18,19,21 is
obtained with selectivities up to 94 % for [Rh(acac)(CO)2]/20 system.
A series of rhodium catalysts modified with phosphorus ligands (23–24) without
perfluoroalkyl substituents have also been evaluated for the hydroformylation of
1-octene using CO2 as the solvent (Scheme 13.13) [46]. It is demonstrated that the
catalyst derived from 24 is very effective at different CO2 densities, and an initial
turnover frequency (TOF) of 3  104 molaldehydemolRhh1 can be obtained.
ScCO2 can also be applied as a reaction medium for asymmetric hydrofor-
mylation [47–49]. High enantioselectivity and regioselectivity are achieved in
rhodium-catalyzed asymmetric hydroformylation with the perfluoroalkyl-
substituted ligand (R,S)-3-H2F6-BINAPHOS in compressed CO2 (Scheme 13.14)
[47]. The isomer (R)-25 is preferentially formed with up to 94 % ee.
Xiao et al. have also examined homogeneous hydroformylation in CO2 [50, 51].
Fast and regioselective hydroformylation of alkyl acrylates can readily be accom-
plished in scCO2 via specific solvent–solute interactions [50]. TOF of up to 1,827 h1
is obtained for the formation of aldehyde by using Rh-P( p-C6H4C6F13)3 as catalyst in
scCO2 while less than 200 h1 in toluene. A fluorous polymeric phosphine could
enhance the solubility of aryl phosphine in scCO2 and thus promotes reaction
(Scheme 13.15) [51]. Unique chemoselectivities are obtained in the hydroformylation
of usually unreactive alkyl acrylates. When equimolar mixture of dec-1-ene and ethyl
acrylate is used as substrate, only a TOF of 872 h1 is measured for ethyl acrylate
converting into the branched aldehyde, while no product arising from dec-1-ene
could be detected.
346 R. Ma et al.

CHO

CHO
Rh cat./L
+
CO/H2
CO2 (d=0.68 g/cm3)
(R)-25
40 °C, 66 h
94% ee
L

(CH2)2(CF2)6F
PAr2
O
O P Ar=
O

Scheme 13.14 Asymmetric hydroformylation in compressed CO2 catalyzed by rhodium catalyst [47]

L:

O O
Et Et CHO O
O [Rh(acac)(CO)2]/L O (TOF 872h-1) O
+ +
CO-H2 (3 MPa) 1 CH2 5
PPh2
C8H17 scCO2, 80 °C, 1 h C8H17 (no reaction) CH2
(CF2)7
CF3

Scheme 13.15 Chemoselective hydroformylation of C═C bonds in scCO2 [51]

[PdCl2(PhCN)2]/L COOR COOR


C6H13 C6H13 + +
CO 3 MPa, ROH C6H13 C6H13
scCO2, 90 °C, 12 h
L: CF3

P CF3 or P
3 3

Scheme 13.16 Hydroesterification of oct-1-ene in scCO2 [52]

13.3.2 Hydroesterification

Masdeu-Bultó has presented the first hydroesterification of linear alkenes in scCO2


using palladium catalysts with phosphines containing –CF3 groups (Scheme 13.16)
[52]. Conversions up to 67 % with selectivities toward the ester up to 64 % can be
obtained by using the catalytic precursor formed from [PdCl2(PhCN)2] and phos-
phine ligands.
13 Homogeneous Catalysis Promoted by Carbon Dioxide 347

13.4 Carbon–Carbon Bond Formation

13.4.1 Diels–Alder Cycloaddition

A mass of literatures exist in relation to Diels–Alder chemistry in supercritical


fluids, especially scCO2. In 1987, Paulaitis and Alexander investigated the earliest
Diels–Alder reaction of maleic anhydride and isoprene in scCO2 (Scheme 13.17)
[53]. Reaction rates are measured over a range of pressures (8–43 MPa) and at three
temperatures (35, 45, and 60  C). At 35  C, the rate constants obtained in scCO2
(above approximately 20 MPa) are similar to those obtained in liquid ethyl acetate.
Ikushima has also studied the Diels–Alder reaction between maleic anhydride
and isoprene catalyzed by aluminum chloride in scCO2 [54]. High-pressure
FT–IR spectroscopy is used to explore this reaction. A modification to the generally
accepted concerted mechanism is proposed that the product is formed in a two-step
sequence of sigma bond formation, followed by cyclization as depicted in
Scheme 13.18.
Regiochemical course of Diels–Alder reactions in scCO2 and conventional
media is also investigated [55]. On analysis of the total reaction mixture, no
deviation from the normal selectivity is observed when performing the reaction
under single-phase conditions (50  C, 9.5 MPa), and similar selectivity is observed
in scCO2 and toluene (Scheme 13.19).

Scheme 13.17 O
Diels–Alder reaction of Me
O O O scCO2 (8–43 MPa)
maleic anhydride and + O
isoprene in scCO2 [53] 35–60 °C Me
O

Scheme 13.18 A two-step O O


mechanism for the
Diels–Alder reaction in AlCl3
+ O O
scCO2 [54]
scCO2, 33 °C
O O

AlCl3

σ O
O
σ
+ O
O

δ O
O
AlCl3
348 R. Ma et al.

Scheme 13.19 CO2Me


Regioselectivity of + +
Diels–Alder reactions in CO2Me CO2Me
scCO2 [55]
PhCH3 145 °C, 15 h 71:29

CO2 (9.5 MPa, 50 °C, 7 d) 71:29

Me3SiO Ph Ph
N Sc(OSO2C8F17)3 (5 mol%) N
+
scCO2 (15 MPa)
Ph H Ph O
50 °C, 24 h
OMe Yield 74%

Scheme 13.20 Lewis acid-catalyzed aza-Diels–Alder reactions in scCO2 [57]

O
I O cat. Pd(OAc)2/26 Ph
26
+ Et3N(1.1equiv) O P
O
scCO2
100 °C, 64 h C6F13 C6F13
Yield 91%

Scheme 13.21 Palladium-catalyzed carbon–carbon bond formation in scCO2 [58]

The deficient rates and selectivities of the uncatalyzed Diels–Alder reaction limit
its synthetic value. In this context, Rayner and co-workers have developed the
Lewis acid-catalyzed Diels–Alder system by Sc(OTf)3 [56]. Quantitative conver-
sion can be achieved with a catalyst loading of 6.5 mol% at 50  C, within 15 h,
whereas the uncatalyzed reaction gives only 10 % complete after 24 h under similar
conditions. Sc(OSO2C8F17)3 can also be used as Lewis acid catalyst for
Diels–Alder reaction and aza-Diels–Alder reaction in scCO2 [57]. 99 % Yield of
the aza-Diels–Alder reaction of Danishefsky’s diene and imine is obtained with
catalyst loading of 5 mol% employed (Scheme 13.20).

13.4.2 Coupling Reaction

Palladium-catalyzed coupling reactions in scCO2 have received much attention.


In 1988, Holmes et al. as well as Tumas et al. have independently developed the
palladium-catalyzed carbon–carbon coupling reactions in scCO2 [58, 59]. The
solubility of the palladium complexes can be dramatically enhanced by introducing
fluorinated phosphine ligand 26. In this regard, the Heck reaction between
iodobenzene and electron-deficient alkenes catalyzed by soluble palladium com-
plexes in scCO2 gives a superior yield (up to 92 %) than that reported in conven-
tional solvents (Scheme 13.21). The workup procedures are significantly easier than
13 Homogeneous Catalysis Promoted by Carbon Dioxide 349

F3C CF3
I 27
Pd(dba)3/27
+ SnBu3
scCO2 (34.5 MPa)
F3C P CF3
90 °C, 5 h Conversion >99%
Selectivity 99%

CF3 CF3

Scheme 13.22 Stille cross-coupling reaction in scCO2 [59]

B(OH)2 Pd(OAc)2-P(t-Bu)3 (5 mol%)


I Me2N(CH2)6NMe2
+
scCO2
120 °C,16 h Yield 76%

Scheme 13.23 Non-fluorous system for coupling reaction in scCO2 [61]

those associated with standard reaction conditions because the extraction of the
product with an organic solvent is no longer needed. Tumas has found that
fluorinated phosphines, particularly tris[3,5-bis(trifluoromethyl)phenyl]phosphine
27, show good activity (86 % conversion) and a quantitative conversion (>99 %)
for Stille cross-coupling reaction between iodobenzene and vinyl(tributyl)tin in
scCO2 (Scheme 13.22).
The nature of the initial palladium source is also crucial. In this context,
fluorinated palladium sources such as Pd(OCOCF3)2 and Pd(F6-acac)2 are superior
catalyst in comparison with the non-fluorinated sources such as Pd(OCOCH3)2 and
Pd2(dba)3 for carrying out a variety of the palladium-catalyzed coupling reactions
in scCO2 with lower catalyst loadings and in shorter reaction times [60]. Although
some phosphine ligands are usually considered to be inferior ligands for coupling
reactions, they also give pleased conversions under this catalytic system. While the
completely non-fluorous Pd(OAc)2–P(t-Bu)3 catalyst system is also highly active to
promote Heck and Suzuki reactions in scCO2 (Scheme 13.23) [61].
Recently, a convenient procedure for the Suzuki–Miyaura cross-coupling has
been developed in scCO2 [62]. The catalyst system is composed of Pd(OCOCF3)2,
Buchwald phosphine ligand, cheap inorganic base such as potassium carbonate, and
a crown ether, giving high yields of the corresponding cross-coupling products.
Markedly, three or even four chlorine atoms in 1,2,4,5-tetrachlorobenzene are
replaced by phenyl groups through treatment with phenylboronic acid (5 equiv.)
in scCO2 in the presence of the developed catalytic system (Scheme 13.24).
Taking a different process, Pd(II) complexes 28 containing various fused bicy-
clic oxadiazoline and tricyclic ketoimine ligands are described as highly efficient
system in the Suzuki–Miyaura reaction in scCO2 [63]. Quantitative cross-coupling
yields are obtained of the aryl halides with phenylboronic acid in short reaction
times (1 h) and very low catalyst loadings (2.5  105 mmol) as shown in
Scheme 13.25.
350 R. Ma et al.

Pd(OCOCF3)2 /L
K2CO3 (6 equiv) PCy2
Cl Cl Ph Ph Ph Ph L:
18-crown-6 (5 mol%) OMe
+
Cl Cl CO2 11 MPa Ph Cl Ph Ph
110 °C, 18 h 41%
59% MeO
Cy=cyclohexyl

Scheme 13.24 Suzuki–Miyaura cross-coupling of tetrachloroarenes in scCO2 [62]

OMe
OMe
cat 28:
cat.28 (2.5x10−5 mmol) Me
+ B(OH)2
scCO2 (12 MPa) Cl2Pd N O
120 °C, 1 h N
I
2
Yield 92%

Scheme 13.25 A Pd(II) complex with bicyclic oxadiazoline for Suzuki–Miyaura reaction [63]

13.4.3 Olefin Metathesis

Supercritical fluids are also useful reaction media for the olefin metathesis which
refers to the mutual alkylidene exchange reaction of alkenes [64, 65]. Transition
metal-catalyzed olefin metathesis reactions in compressed CO2 media have
been developed by Leitner and co-workers [66]. Ring-opening metathesis poly-
merization (ROMP) of norbornene and cyclooctene gives the corresponding poly-
mer in excellent yields by using the conventional metathesis catalysts 29 and 30
(Scheme 13.26), both in liquid CO2 and scCO2. This methodology for ring-closing
metathesis (RCM) is also remarkable by using the same carbene complexes.
A remarkable influence of the density of the supercritical medium on the product
of RCM of 31 is observed with the 16-membered ring 32 being formed in excellent
yield at CO2 density of 0.65 g · ml1, oligomers are obtained by acyclic diene
metathesis (ADMET) at lower CO2 density (d < 0.65 g · ml1) as depicted in
Scheme 13.26.
Another remarkable feature is increasing the scope of this versatile transformation
and applying scCO2 as a protective medium. The epilachnene 34 has been synthe-
sized by RCM of diene 33 without the need for N-protection (Scheme 13.27) [67].

13.4.4 Aldol Reaction

Ikariya has disclosed the first catalytic asymmetric Mukaiyama aldol reaction by a
chiral binaphthol–titanium(IV) catalyst in scCO2, although, giving a very low yield
13 Homogeneous Catalysis Promoted by Carbon Dioxide 351

PCy3 Ph
Cl
Ru
Ph
Cl H3C(F3C)2CO Mo
PCy3 Ph
H3C(F3C)2CO
29 30
O

O 29
cat
Ru
O2
O scC L
5 g/m RCM product
d >0.6
32

Ru
cat
29
scC
d<0 O
31 .65 2
g/m oligomers
L ADMET products

Scheme 13.26 Ring-closing metathesis catalyzed by carbene complexes in scCO2 [66]

O O

O Ru cat 29 (1 mol%) O
NH scCO2 NH
40 °C, 72 h
Yield 74%
33 34

Scheme 13.27 Synthesized epilachnene 34 by RCM of diene 33 in scCO2 [67]

of the aldol product [68]. Poly(ethylene glycol) (PEG) derivatives have remarkable
effect on Mannich and aldol reactions in scCO2 [69, 70]. It is envisaged that
utilization of PEG as surfactants media can accelerate the reactions by forming
emulsions in a single scCO2 phase. Indeed, PEG400 is found to be effective in the
Yb(OTf)3-catalyzed Mannich reactions of imine with silyl enolate, giving rise to
72 % yield of product, while only 10 % yield of the product is obtained without
PEG (Scheme 13.28). In the case of the Sc(OTf)3-catalyzed aldol reactions of silyl
enolates with aldehydes, the poly(ethylene glycol) dimethyl ether (PEG(OMe)2,
average Mw ¼ 500) is more efficient than PEG itself (Scheme 13.29).
352 R. Ma et al.

Bn Yb(OTf)3 (5 mol%) Bn
N OSiMe3 NH O
+ PEG400
Ph H Ph OMe
OMe CO2 15 MPa
50 °C, 3 h

Yield 72%

Scheme 13.28 Mannich reaction catalyzed by Yb(OTf)3 in scCO2 [69]

O OH O
OSiButMe2 Sc(OTf)3 (5 mol%)
H PEG(OMe)2 OEt
+
OEt CO2 8 MPa
50 °C, 3 h
Yield 89%

Scheme 13.29 Aldol reaction in CO2–PEG(OMe)2 system [69]

PdCl2[P(OEt)3]2 O
I Et3N (2.2 equiv.)
+ CO O
OH scCO2
0.1 MPa 130 °C, 18 h
TON=4650

Scheme 13.30 Pd complex-catalyzed intramolecular carbonylation in scCO2 [71]

13.4.5 Miscellaneous Reactions

Pd complexes with non-fluorous ligands have been developed as catalyst for the
intramolecular carbonylation reactions of aryl halides in scCO2 [71]. Turnover
number (TON) of 1,880 is obtained when PdCl2(MeCN)2 is used as a catalyst in
scCO2 for the intramolecular carbonylation of 2-iodobenzyl alcohol. No marked
effect of CO pressure on the TON values is observed in scCO2, because of the
higher diffusivity of scCO2 as well as the high solubility of CO in this medium.
Near completion (TON ¼ 4,650) of the carbonylation can be reached even pro-
ceeding at low CO pressure (0.1 MPa) with the more soluble triethylphosphite as a
ligand (Scheme 13.30). The rate of the intramolecular carbonylation reaction is
found to be higher than that in toluene.
The Pauson–Khand reaction, the catalytic co-cyclization of an alkyne with an
alkene and CO, leading to cyclopentanone, has been recognized as one of the most
optimal routes to the synthesis of cyclopentenones [72, 73]. Jeong et al. have
reported several examples of the intramolecular Pauson–Khand reaction catalyzed
by Co2(CO)8 in scCO2 [74]. Reaction of the intramolecular enynes and CO is
optimized to give 85 % yield of cyclopentanones with 2.5 mmol% catalyst loading
under a total pressure of 18 MPa (Scheme 13.31). The reaction also proceeds pretty
13 Homogeneous Catalysis Promoted by Carbon Dioxide 353

EtO2C
EtO2C 2.5 mol% Co2(CO)8 EtO2C
+ CO
scCO2(total18 MPa) EtO2C O
1.5 MPa 69 °C
Yield 85%

Scheme 13.31 Intramolecular Pauson–Khand reaction in scCO2 [74]

O H
3 mol% Co2(CO)8
+ + CO Ph
scCO2(total18 MPa)
88 °C
H
Yield 87%

Scheme 13.32 Intermolecular Pauson–Khand reaction in scCO2 [74]

O2N
O H
OH
CO2Me
CO2Me CO2Me
+ DABCO O
CO2 (12 MPa) O N -H2O CO2Me
2
NO2 50 °C, 24 h Yield 64%
O2N
35 36

Scheme 13.33 DABCO catalyzed the Baylis–Hillman reaction in scCO2 [77]

for a number of substituted enynes. Intermolecular coupling reaction between


phenyl acetylene and norbornadiene also runs well, giving the bicyclic compound
in 87 % yield (Scheme 13.32).
The Baylis–Hillman reaction is considered to be the useful carbon–carbon bond
forming reaction between activated alkenes and aldehydes, introducing a
hydroxyalkyl moiety at the α-position of Michael acceptors [75, 76]. The reaction
between acrylate and a variety of aromatic aldehydes can be efficiently carried out
in scCO2 by using DABCO (1,4-diazabicyclo[2.2.2]octane) as a catalyst [77].
Enhanced reaction rates comparable to the solution-phase reactions are observed
under this catalyst system. It is noted that an unprecedented dimerization 36 is
observed as dimers of the original Baylis–Hillman adducts 35 at low pressure
(Scheme 13.33).
Palladium-catalyzed cyclotrimerization of different substituted alkynes occurs
smoothly in good yield and regioselectivity in scCO2 (Scheme 13.34) [78]. CuCl2,
as a co-catalyst, could play an important role in enhancing the reaction rate
and regenerating the active PdCl2. Addition of methanol can increase the solubility
of PdCl2 and CuCl2 in scCO2. Also, environmentally friendly scCO2 may be a
substitute for toxic organic solvents and show potential utility in industry. CpCo
(CO)2 also has been reported as a catalyst for the cyclotrimerization of
monosubstituted and disubstituted alkynes in scCO2 [79].
354 R. Ma et al.

Ph
PdCl2/CuCl2 H3C CH3
Ph CH3
CO2 16 MPa, MeOH 5 mL Yield 92%
Ph Ph
40 °C, 8 h
CH3

Scheme 13.34 Palladium-catalyzed cyclotrimerization in scCO2 [78]

13.5 Oxidation Reaction

Since CO2 is the highest oxidation state of carbon, its further oxidation will not
proceed. Therefore, by-products originating from CO2 can be depressed by using
compressed CO2 as reaction media. On the other hand, compressed CO2 provides a
safe reaction environment for reactions due to its excellent mass and heat transfer
properties. Dense CO2 appears to be a unique solvent for oxidation reactions
[80, 81]. Moreover, the high miscibility of the oxidants, especially gaseous O2, in
scCO2 also can eliminate interphase transport limitations [5, 82].

13.5.1 Oxidation Reaction of Alcohols

Selective aerobic oxidation of alcohols to aldehydes or ketones has been widely


studied not only in conventional liquid solvents but also in scCO2. The key to
achieve high selectivity of aldehydes is to suppress the formation of carboxylic
acid, which can further react with reactant alcohols to yield esters, leading to a
further decline of the aldehyde selectivity [10].
The copper-based catalyst with a small amount of cosolvent (toluene or
fluorobenzene) is successfully developed for the oxidation of cyclohexanol to
cyclohexanone in compressed CO2 [83]. The results indicate that the selectivity
of the reaction in compressed CO2 with or without cosolvent is much higher than
that in the liquid solvents or without solvent (Table 13.1). Cosolvent could enhance
the reaction rate by the means of tuning the properties of scCO2 and the kinetic and
thermodynamic properties of reactions.
The copper acetate complex 38, i.e., the polydimethylsiloxane-functionalized
pyridine 37, is developed for the aerobic oxidation of alcohols to the correspond-
ing aldehydes in scCO2 as depicted in Scheme 13.35 [84]. In combination with
2,2,6,6-tetramethylpiperidin-1-yloxy (TEMPO) as co-catalyst, complete conver-
sion of 4-nitrobenzyl alcohol could be achieved at 15 MPa CO2, 1 MPa O2, and
60  C within 4 h (Scheme 13.36). However, the analogous copper derivatives
containing simpler pyridine substituents, [Cu(AcO)2(py)]2 and [Cu(AcO)2(4VP)]2
(4VP ¼ 4-vinylpyridine), show similar activities in spite of the negligible solubil-
ity in scCO2. In addition, the solubility of the metal complex is unnecessary for the
oxidation of alcohols.
13 Homogeneous Catalysis Promoted by Carbon Dioxide 355

Table 13.1 Solvent effect on the aerobic oxidation of cyclohexanol

OH + 1/2 O2 O + H2O
60⬚C, 6 h

Solvent Conversion of alcohol (%) Selectivity of product (%)


No solvent 19.8 70.6
Toluene 33.4 77.9
Fluorobenzene 26.5 75.2
CO2a 7.51 98.9
CO2a,b 30.0 89.4
CO2+toluenea,c 18.6 97.3
a
Apparent density, 0.77 g/mL
b
Reaction time, 36 h
c
The amount of cosolvent in CO2 is 0.5 mol%

Si H Si C2H4 N 37
O O
n n

[Cu(AcO)2(4VP)]2 [Cu(AcO)2(H2O)]2

Si C2H4 N [Cu] 38
O n

Scheme 13.35 Synthesis of compounds 37 and 38 [84]

[Cu]/TEMPO/O2
O2N CH2OH O2N CHO
scCO2

Scheme 13.36 Alcohol oxidation catalyzed by Cu-TEMPO [84]

13.5.2 CO2 as a Soft Oxidant for the Oxidation of Aldehydes

The utilization of CO2 as a soft oxidant is a promising concept for industrial


applications, such as the oxidative coupling of CH4 and the oxidative dehydroge-
nation of alkanes and alkyl aromatics [85]. On the other hand, aromatic aldehydes
can also be oxidized to the corresponding carboxylic acids under mild conditions
using CO2 as a soft oxidant.
The catalytic oxidation of aromatic aldehydes under mild conditions is developed
by using CO2 as oxidant and N-heterocyclic carbenes (NHCs) as organocatalysts
(Scheme 13.37) [86]. Under the optimized conditions, various aldehydes are oxidized
to corresponding acids with 10 mol% 1,3-dimesitylimidazol-2-ylidene (IMes) as
organocatalyst at room temperature.
356 R. Ma et al.

Scheme 13.37 Aldehyde CO2


oxidation reaction with CO2
catalyzed by NHC [86] O NHC/base O

R H polar solvent R OH

R= Aromatic group CO

CO2
N
N N
N
O
O
39
O

CO

N H KO O K
N N O
N
O
O
O O

41 O K 40

Scheme 13.38 Catalytic cycle in oxidation of cinnamaldehyde by NHC [86]

The mechanism of direct activation of CO2 by the NHC is proposed as depicted


in Scheme 13.38. Firstly, NHC reacts with CO2 resulting in imidazolium carbox-
ylate 39, which attacks the 2-position of cinnamaldehyde, generating the possible
intermediate 40. And then, the hydrogen from the 2-position carbon on
cinnamaldehyde is trapped by the nitrogen on imidazolium carboxylate, forming
a stable six-membered ring intermediate. Finally, CO is released from NHC–CO
complex 41, and NHC is reused for the next cycle.
NHC-mediated oxidation of aldehydes could also proceed via the addition of CO2
to the readily formed Breslow intermediate as illuminated in Scheme 13.39 [87]. The
reaction implements via the addition of CO2 to the readily formed Breslow interme-
diate 42 to form the hydroxycarboxylate 43 and the tautomer 44, which could lose
CO and hydroxide to afford the benzoic acid as depicted in Scheme 13.40.
Support for this hypothesis is obtained by the detection of CO in the reaction
mixture. Indirect evidence for the mechanistic postulate outlined above is accrued
by the formation of benzoic acid when phenylglyoxylic acid is exposed to NHC
13 Homogeneous Catalysis Promoted by Carbon Dioxide 357

Cl
Mes N N Mes
Ar CHO + CO2 Ar COOH
DBU, THF

Scheme 13.39 Oxidation of aromatic aldehydes to carboxylic acids with CO2 [87]

Mes RCHO Mes Mes


N O
N N OH
O =C =O
N O
N N R R
Mes OH
Mes Mes
42 43

O
Mes
R OH Mes
N O N O
-CO
N R N OH
-OH
Mes OH Mes R O

44

Scheme 13.40 Proposed mechanistic pathway for the oxidation of aldehydes through Breslow
intermediate [87]

PdCl2 + CO + H2O Pd + 2HCl + CO2

Mes
N
O Cl O
N
OH
Mes OH
O DBU, THF, rt, 16 h
62%

Scheme 13.41 Evidence for NHC-mediated oxidation of aldehydes [87]

under the same experimental conditions in which benzaldehyde is transformed to


benzoic acid as illustrated in Scheme 13.41.
However, there is debate about the oxidation ability of CO2 for the transforma-
tion of cinnamic aldehydes to acids catalyzed by NHCs. A study on NHC-catalyzed
oxidation of aldehydes by using CO2 as the stoichiometric oxidant is conducted,
358 R. Ma et al.

Scheme 13.42 Oxidation [Fe(TPP)Cl], t-BuOOH


of cyclohexane catalyzed by OH
scCO2
[Fe(TPP)Cl] in scCO2 [90]

[Fe(TPP)Cl]:

N N
Fe
N N

concluding that the apparent NHC-catalyzed oxidation of aldehydes with CO2


actually involves air as the stoichiometric oxidant [88]. Series of computational
and experimental investigation on the role of CO2 in NHC-catalyzed oxidation of
aldehydes is also carried out, indicating that this reaction could not be performed at
room temperature and without the assistance of H2O [89].

13.5.3 Oxidation Reaction of Hydrocarbons

Oxidation of cyclohexane could be achieved in scCO2 catalyzed by iron tetraphe-


nylporphyrin ([Fe(TPP)Cl]) with tert-butyl-hydroperoxide (t-BuOOH) as terminal
oxidants (Scheme 13.42) [90]. CO2 density is an important variable for the selec-
tivity in the oxidation process. The selectivity is higher in scCO2 than that in
organic solvents with up to 15 % yield of cyclohexanol.
The aerobic oxidation of cyclohexene could be achieved by Fe(III)(5,10,15,20-
tetrakis(pentafluorophenyl)porphyrin)Cl catalyzed at 50  C, 16.5 MPa of the total
pressure with 2 MPa O2 partial pressure [91]. The reaction is retarded in the
presence of the radical inhibitor 2,6-di-tert-butyl-4-methylphenol, suggesting that
the product is formed through free radical allylic oxidation mechanism.

13.5.4 Oxidation Reaction of Olefins

An alternative product of oxidation of olefins in CO2 medium is epoxide.


Cyclohexene is epoxidized to 1,2-cyclohexene oxide and 1,2-cyclohexanediol by
using scCO2 as both solvent and reactant in combination with aqueous H2O2, which
is made possible through the in situ formation of peroxycarbonic acid [92].
By adding NaHCO3 and the hydrophilic cosolvent N, N-dimethylformamide
(DMF), cyclohexene conversion is increased from 0.4 to 12.6 mol%.
13 Homogeneous Catalysis Promoted by Carbon Dioxide 359

Scheme 13.43 The R2OH HCl


proposed mechanism for
the Wacker reaction in
ROH/scCO2 [93]
R1CH CH2
OR2
PdCl2L
R1CHCH2PdClL
L
L CuCl2
OR2 O
HCl
R1C CH2 R1CCH3
R1CH CH2
PdCl2L2
CuCl

MeO Lower CO2 pressure


CO2Me Higher O pressure
2
OMe Lower reaction temperature
+ up to 99.7% yield
PdCl2
CO2Me + O2 +MeOH MeO
scCO2 CO2Me
+
CO2Me
Higher CO2 pressure
Lower O2 pressure
Higher reaction temperature
MeO2C CO2Me up to 69.3% yield

Scheme 13.44 PdCl2-catalyzed aerobic oxidation of terminal alkenes with methanol in


scCO2 [94]

Wacker reaction is one of the most important reactions for the functionalization
of alkene feedstock. The Wacker reaction could be carried out smoothly in scCO2
or ROH/scCO2 [93]. In order to increase the solubility of catalysts and accelerate
the reaction rate, a suitable amount of MeOH/EtOH is added to modify the solvent
polarity of scCO2. The reaction in scCO2 or MeOH/scCO2 is homogeneous since
organopalladium complexes are formed during the oxidative process as illuminated
in Scheme 13.43.
Acetalization of terminal alkenes with alcohols, which undergo a Wacker-type
process, is another important reaction for the functionalization of alkene feedstock.
The PdCl2-catalyzed aerobic oxidation of terminal alkenes with electron-
withdrawing groups is performed in scCO2 (Scheme 13.44). The PdCl2-catalyzed
aerobic oxidation of terminal alkenes with electron-withdrawing groups runs in
scCO2 through optimizing reaction conditions, to yield acetals and benzene deriv-
atives (Scheme 13.44) [94]. Higher CO2 pressure, lower O2 pressure, and higher
reaction temperature could favor the formation of the trisubstituted benzene,
whereas the opposite tendency is observed for the generation of the dimethoxy
adduct.
360 R. Ma et al.

13.5.5 Baeyer–Villiger Reaction

The Baeyer–Villiger oxidation of ketones into esters or lactones is one of the


most widely applied transformations in organic synthesis [95]. The aerobic
Baeyer–Villiger oxidation can be performed very efficiently in compressed CO2
at room temperature using oxygen as primary oxidant and an aldehyde as
co-reductant [96]. No additional catalyst is required, and good to excellent yields
are achieved for a wide range of cyclic and acyclic ketones.

13.6 Polymerization

13.6.1 Free Radical Polymerization in ScCO2

Supercritical fluids can also provide insight into the general roles that solvents play
in polymerization reactions. In 1992, the successful synthesis of high molar mass
fluoropolymers was covered by using homogeneous free radical polymerization
methods in scCO2 [97]. Azobisisobutyronitrile (AIBN) is found to be a more
efficient radical generator in scCO2 than in benzene while with a slower decompo-
sition rate, due to the absence of a cage effect in the low viscosity of CO2. The
perfluoropolymer 46 made in scCO2 (59.4  C, 207 bar) through the homopoly-
merization of 1,1-dihydroperfluorooctyl acrylate (FOA) 45 has a molar weight of
270,000 g · mol1 with 65 % yield (Scheme 13.45).

13.6.2 Cationic Polymerization in ScCO2

The choice of solvent plays an important role in controlling the course of cationic
polymerizations. The utilization of liquid and supercritical CO2 can perform
as a solvent-dispersing medium for the cationic polymerization of vinyl ethers
and oxetanes [98]. Interestingly, polymerization of the hydrocarbon vinyl ether
47 catalyzed by EtAlCl2 and initiated by isobutyl vinyl ether adduct 50 is initially
homogenous in scCO2 and then becomes heterogeneous as the polymer 49

O n AIBN:
AIBN CH3 CH3
O C7F15 O O
scCO2 H3C N N CH3
59.4 °C, 20 MPa C7F15 CN CN
45 46 Yield: 65%

Scheme 13.45 Homopolymerization of 1,1-dihydroperfluorooctyl acrylate in scCO2 [97]


13 Homogeneous Catalysis Promoted by Carbon Dioxide 361

1) EtAlCl2 /ethyl acetate


scCO2 OR
O
30–60 °C H
OR + OEt
O O 2) sodium ethoxide n
i-Bu
50 49
47 R=CH2CH(CH3)2
48 R=(CH2)2N(C3H7)SO2C8F17

Scheme 13.46 Polymerization of vinyl ethers EtAlCl2 in scCO2 [98]

O
O
O Sn(Oct)2/BuOH
O
scCO2 BuO n H
65-80 °C

The active initiating alkoxide species

BuOH
Oct Sn Oct Oct Sn OBu
-OctH

Oct = O BuOH -OctH


O

BuO Sn OBu

Scheme 13.47 Sn(Oct)2-catalyzed ring-opening polymerization of ε-CL in scCO2 [100]

forms and precipitates (Scheme 13.46). Polymers with high molar mass and broad
molecular weight distributions are obtained in the absence of initiator 50.
Fluorocarbon-based alkyl vinyl ether 48 in the absence of initiator 50 is also
polymerized in CO2, and the polymerization remains homogeneous throughout
the course of the reaction with a narrow molecular weight distribution.

13.6.3 Metal-Catalyzed Polymerization in ScCO2

Ring-opening polymerization (ROP) is one of the most facile routes to obtain a


number of linear aliphatic polyesters [99]. The ROP of ε-caprolactone (ε-CL)
catalyzed by the industrially important catalyst Sn(Oct)2 has been developed in
scCO2 [100]. The active initiating alkoxide species in the polymerization is a tin
alkoxide species formed by Sn(Oct)2 and an alcohol as illustrated in Scheme 13.47.
The molecular weight of the product is controlled by the ratio of monomer to
initiator, i.e., the [CL]/[BuOH] ratio, and increasing the ratio leads to an increase in
molecular weight. Increasing the temperature results in an increase in the rate of
362 R. Ma et al.

catalyst 29 or 30
scCO2 + THF
m n

Scheme 13.48 ROMP of norbornene by metal carbene catalysts in scCO2 [102]

Ph

Ph Ph
Ph
cat Rh or Rh/52
Ph +
quinuclidine n
compressed CO2 Ph Ph
Ph n
Ph

(CH2)2(CF2)6F 51 52
53:

F(F2C)6(H2C)2 (CH2)2(CF2)6F

Scheme 13.49 Rh-catalyzed phenylacetylene polymerization in compressed CO2 [103]

polymerization, while increasing the initial pressure of the reaction medium has the
opposite effect and leads to a decrease in rate. The rate of ROP is generally slower
in scCO2 compared to bulk or solution polymerization conducted in toluene.
Ru(H2O)6(Tos)2 has been reported as a catalyst for the ring-opened metathesis
polymers (ROMP) of norbornene in liquid CO2, leading to formation of high cis,
syndiotactic poly(norbornene) [101]. Notably, the former polymers of bulk monomer
in pure CO2 are more cis stereospecific than those in methanol or CO2/methanol and
have a blocky distribution of cis and trans double bonds. Subsequently, the metal
carbene catalysts (29 and 30, Scheme 13.26) have been developed for the ROMP of
norbornene in scCO2 (Scheme 13.48) [102]. Polynorbornenes prepared in scCO2 are
similar to that made in the THF in molecular weight, molecular weight distribution,
and thermal stability. Neither the reaction temperature nor pressure makes significant
differences in the trans/cis ratio of the polynorbornenes obtained in scCO2, while
higher temperature gives higher molecular weights.
The polymerization of phenylacetylene catalyzed by rhodium complex also has
been performed in compressed CO2 [103]. The unmodified catalyst, (norborna-
2,5-diene)rhodium acetylacetanoate [Rh(nbd)(acac)], is insoluble in compressed
CO2. However, the polymerization in CO2 proceeds more efficiently than that in
conventional solvents such as THF or hexane, resulting in the cis-cisoidal product
51 in a 65 % yield. Using the fluorinated phosphine [4-F(CF2)6(CH2)2C6H4]3P 53,
modified rhodium catalyst gives rise to an increase in selectivity toward the
cis-transoidal polymer 52 as shown in Scheme 13.49.
13 Homogeneous Catalysis Promoted by Carbon Dioxide 363

O
[Pd(CH3)(NCCH3)(54 or 55)]BARF
(0.0125 mmol)
+ CO n
O
1-5 bar scCO2

C8F17(CH2)4 C8F17(CH2)4
CF3
N N
B
N 4
N
CF3
C8F17(CH2)4 C8F17(CH2)4
54 55 BARF

Scheme 13.50 CO/tert-butylstyrene copolymerization by palladium cationic complexes in


scCO2 [104]

Polyketone has been formed from the copolymerization of CO and tert-


butylstyrene in scCO2 as a reaction medium, avoiding the use of classical organic
or fluorinated solvents as illustrated in Scheme 13.50 [104]. Palladium cationic
complexes of general formula [Pd(CH3)(NCCH3)(L)]BARF containing mono-
chelated bipyridine 54 or phenanthroline ligands 55 with perfluorinated chains are
used as catalysts in scCO2 for this kind of copolymerization and give rise to
the polyketone of highly syndiotactic and controllable molecular weights.

References

1. Welton T (1999) Room-temperature ionic liquids. Solvents for synthesis and catalysis. Chem
Rev 99(8):2071–2084
2. Lindström UM (2002) Stereoselective organic reactions in water. Chem Rev 102
(8):2751–2772
3. Singh MS, Chowdhury S (2012) Recent developments in solvent-free multicomponent
reactions: a perfect synergy for eco-compatible organic synthesis. RSC Adv 2
(11):4547–4592
4. Jessop PG, Ikariya T, Noyori R (1995) Homogeneous catalysis in supercritical fluids. Science
269(5227):1065–1069
5. Jessop PG, Ikariya T, Noyori R (1999) Homogeneous catalysis in supercritical fluids. Chem
Rev 99(2):475–494
364 R. Ma et al.

6. Beckman EJ (2004) Supercritical and near-critical CO2 in green chemical synthesis and
processing. J Supercrit Fluids 28(2–3):121–191
7. Jessop PG (2006) Homogeneous catalysis using supercritical fluids: recent trends and sys-
tems studied. J Supercrit Fluids 38(2):211–231
8. Page SH, Goates SR, Lee ML (1991) Methanol/CO2 phase behavior in supercritical fluid
chromatography and extraction. J Supercrit Fluids 4(2):109–117
9. Woods HM, Silva MMCG, Nouvel C et al (2004) Materials processing in supercritical carbon
dioxide: surfactants, polymers and biomaterials. J Mater Chem 14(11):1663–1678
10. Seki T, Baiker A (2009) Catalytic oxidations in dense carbon dioxide. Chem Rev
109(6):2409–2454
11. Darr JA, Poliakoff M (1999) New directions in inorganic and metal-organic coordination
chemistry in supercritical fluids. Chem Rev 99(2):495–542
12. Beckman EJ (2003) Green chemical processing using CO2. Ind Eng Chem Res
42(8):1598–1602
13. Lyubimov SE, Tyutyunov AA, Kalinin VN et al (2007) Carboranylphosphites—new
effective ligands for rhodium-catalyzed asymmetric hydrogenation of dimethyl itaconate.
Tetrahedron Lett 48(46):8217–8219
14. Lyubimov SE, Said-Galiev EE, Khokhlov AR et al (2008) The use of monodentate phos-
phites and phosphoramidites as effective ligands for Rh-catalyzed asymmetric hydrogenation
in supercritical carbon dioxide. J Supercrit Fluids 45(1):70–73
15. Lyubimov SE, Davankov VA, Said-Galiev EE et al (2008) Chiral phosphoramidites as
inexpensive and efficient ligands for Rh-catalyzed asymmetric olefin-hydrogenation in
supercritical carbon dioxide. Catal Commun 9(9):1851–1852
16. Lyubimov SE, Kuchurov IV, Davankov VA et al (2009) Synthesis of chiral amino acid
derivatives in supercritical carbon dioxide using Rh-PipPhos catalyst. J Supercrit Fluids
50(2):118–120
17. Lyubimov SE, Kuchurov IV, Tyutyunov AA et al (2010) The use of new carboranylphosphite
ligands in the asymmetric Rh-catalyzed hydrogenation. Catal Commun 11(5):419–421
18. Jessop PG, Hsiao Y, Ikariya T et al (1994) Catalytic production of dimethylformamide from
supercritical carbon dioxide. J Am Chem Soc 116(19):8851–8852
19. Munshi P, Heldebrant DJ, McKoon EP et al (2003) Formanilide and carbanilide from aniline
and carbon dioxide. Tetrahedron Lett 44(13):2725–2727
20. Krocher O, Koppel RA, Baiker A (1997) Highly active ruthenium complexes with bidentate
phosphine ligands for the solvent-free catalytic synthesis of N,N-dimethylformamide and
methyl formate. Chem Commun 1997, (5):453–454
21. Ting SST, Tomasko DL, Foster NR et al (1993) Solubility of naproxen in supercritical carbon
dioxide with and without cosolvents. Ind Eng Chem Res 32(7):1471–1481
22. Xiao J, Nefkens SCA, Jessop PG et al (1996) Asymmetric hydrogenation of α, β-unsaturated
carboxylic acids in supercritical carbon dioxide. Tetrahedron Lett 37(16):2813–2816
23. Berthod M, Mignani G, Lemaire M (2004) New perfluoroalkylated BINAP usable as a ligand
in homogeneous and supercritical carbon dioxide asymmetric hydrogenation. Tetrahedron
Asymmetry 15(7):1121–1126
24. Johansson A (1995) Methods for the asymmetric preparation of amines. Contemp Org Synth
2(6):393–407
25. Kainz S, Brinkmann A, Leitner W et al (1999) Iridium-catalyzed enantioselective hydroge-
nation of imines in supercritical carbon dioxide. J Am Chem Soc 121(27):6421–6429
26. Lyubimov SE, Rastorguev EA, Petrovskii PV et al (2011) Iridium-catalyzed asymmetric
hydrogenation of imines in supercritical carbon dioxide using phosphite-type ligands. Tetra-
hedron Lett 52(12):1395–1397
27. Wang S, Kienzle F (2000) The syntheses of pharmaceutical intermediates in supercritical
fluids. Ind Eng Chem Res 39(12):4487–4490
13 Homogeneous Catalysis Promoted by Carbon Dioxide 365

28. Turova OV, Kuchurov IV, Starodubtseva EV et al (2012) Ru–BINAP-catalyzed asymmetric


hydrogenation of keto esters in high pressure carbon dioxide. Mendeleev Commun
22(4):184–186
29. Kainz S, Koch D, Leitner W et al (1997) Perfluoroalkyl-substituted arylphosphanes as ligands
for homogenous catalysis in supercritical carbon dioxide. Angew Chem Int Ed
36(15):1628–1630
30. Guzel B, Omary MA, Fackler JP Jr et al (2001) Synthesis and characterization of {[(COD)Rh
(bis-(2R,3R)-2,5-diethylphospholanobenzene)]+BARF} for use in homogeneous catalysis
in supercritical carbon dioxide. Inorg Chim Acta 325(1–2):45–50
31. Horishi Nishida NT, Yoshimura M, Sonoda T, Kobyrashi H (1984) Tetrakis[3,5-bis
(trifluoromethyl)phenyl]borate. Highly lipophilic stable anionic agent for solvent-extraction
of cations. Bull Chem Soc Jpn 57(9):2600–2604
32. Altinel H, Avsar G, Guzel B (2009) Fluorinated rhodium-phosphine complexes as efficient
homogeneous catalysts for the hydrogenation of styrene in supercritical carbon dioxide.
Transit Metal Chem 34(3):331–335
33. Burkhardt ER, Matos K (2006) Boron reagents in process chemistry: excellent tools for
selective reductions. Chem Rev 106(7):2617–2650
34. Zhu Y, Jang SHA, Tham YH et al (2011) An efficient and recyclable catalytic system
comprising nano-Iridium(0) and a pyridinium salt of nido-carboranyldiphosphine for the
synthesis of one-dimensional boronate esters via hydroboration reaction. Organometallics
31(7):2589–2596
35. Carter CAG, Baker RT, Tumas W et al (2000) Enhanced regioselectivity of rhodium-catalysed
alkene hydroboration in supercritical carbon dioxide. Chem Commun 2000, (5):347–348
36. Bhattacharyya P, Gudmunsen D, Hope EG et al (1997) Phosphorus(III) ligands with fluorous
ponytails. J Chem Soc Perkin Trans 1(24):3609–3612
37. Rathke JW, Klingler RJ, Krause TR (1991) Propylene hydroformylation in supercritical
carbon dioxide. Organometallics 10(5):1350–1355
38. Guo Y, Akgerman A (1997) Hydroformylation of propylene in supercritical carbon dioxide.
Ind Eng Chem Res 36(11):4581–4585
39. Jessop PG, Ikariya T, Noyori R (1995) Selectivity for hydrogenation or hydroformylation of
olefins by hydridopentacarbonylmanganese(I) in supercritical carbon dioxide. Organometal-
lics 14(3):1510–1513
40. Bach I (1998) Hydroformylation of hex-1-ene in supercritical carbon dioxide catalysed by
rhodium trialkylphosphine complexes. Chem Commun 1998, (14):1463–1464
41. Sellin MF, Cole-Hamilton DJ (2000) Hydroformylation reactions in supercritical carbon
dioxide using insoluble metal complexes. J Chem Soc Dalton Trans (11):1681–1683
42. Palo DR, Erkey C (1998) Homogeneous catalytic hydroformylation of 1-octene in supercrit-
ical carbon dioxide using a novel Rhodium catalyst with fluorinated arylphosphine ligands.
Ind Eng Chem Res 37(10):4203–4206
43. Patcas F, Maniut C, Ionescu C et al (2007) Supercritical carbon dioxide as an alternative
reaction medium for hydroformylation with integrated catalyst recycling. Appl Catal B
Environ 70(1–4):630–636
44. Van Rooy A, de Bruijn JNH, Roobeek KF et al (1996) Rhodium-catalysed hydroformylation
of branched 1-alkenes; bulky phosphite vs. triphenylphosphine as modifying ligand.
J Organomet Chem 507(1–2):69–73
45. Estorach CT, Orejón A, Masdeu-Bultó AM (2008) New rhodium catalytic systems with
trifluoromethyl phosphite derivatives for the hydroformylation of 1-octene in supercritical
carbon dioxide. Green Chem 10(5):545
46. Koeken ACJ, Benes NE, van den Broeke LJP et al (2009) Efficient hydroformylation in dense
carbon dioxide using phosphorus ligands without perfluoroalkyl substituents. Adv Synth
Catal 351(9):1442–1450
47. Francio G, Leitner W (1999) Highly regio- and enantio-selective rhodium-catalysed asym-
metric hydroformylation without organic solvents. Chem Commun 1999, (17):1663–1664
366 R. Ma et al.

48. Franciò G, Wittmann K, Leitner W (2001) Highly efficient enantioselective catalysis in


supercritical carbon dioxide using the perfluoroalkyl-substituted ligand (R, S)-3–H2F6-
BINAPHOS. J Organomet Chem 621(1–2):130–142
49. Kainz S, Leitner W (1998) Catalytic asymmetric hydroformylation in the presence of
compressed carbon dioxide. Catal Lett 55(3–4):223–225
50. Hu Y, Chen W, Osuna AMB et al (2001) Rapid hydroformylation of alkyl acrylates in
supercritical CO. Chem Commun 2001, (8):725–726
51. Hu Y, Chen W, Osuna AMB et al (2002) Fast and unprecedented chemoselective hydrofor-
mylation of acrylates with a fluoropolymer ligand in supercritical CO2. Chem Commun
2(7):788–789
52. Estorach CT, Masdeu-Bultó AM (2007) Hydroesterification of 1-alkenes in Supercritical
Carbon dioxide. Catal Lett 122(1–2):76–79
53. Paulaitis ME, Alexander GC (1987) A case study of the thermodynamic solvent effects on
a Diels-Alder reaction in supercritical carbon dioxide. Reactions in supercritical fluids.
Pure Appl Chem 59:61–68
54. Ikushima Y, Saito N, Arai M (1991) High-pressure Fourier transform infrared spectroscopy
study of the Diels-Alder reaction of isoprene and maleic anhydride in supercritical carbon
dioxide. Bull Chem Soc Jpn 64(1):282–284
55. Renslo AR, Weinstein RD, Tester JW et al (1997) Concerning the regiochemical course of
the Diels–Alder reaction in supercritical carbon dioxide. J Org Chem 62(13):4530–4533
56. Scott Oakes R, J. Heppenstall T, Shezad N et al (1999) Use of scandium tris(trifluorometha-
nesulfonate) as a Lewis acid catalyst in supercritical carbon dioxide: efficient Diels-Alder
reactions and pressure dependent enhancement of endo:exo stereoselectivity. Chem Commun
1999, (16):1459–1460
57. Matsuo J-I, Tsuchiya T, Odashima K et al (2000) Lewis acid catalysis in supercritical carbon
dioxide. Use of Scandium tris(heptadecafluorooctanesulfonate) as a Lewis acid catalyst in
Diels-Alder and aza Diels-Alder reactions. Chem Lett 29(2):178–179
58. Carroll MA, Holmes AB (1998) Palladium-catalysed carbon–carbon bond formation in
supercritical carbon dioxide. Chem Commun 1998, (13):1395–1396
59. Morita DK, David SA, Tumas W et al (1998) Palladium-catalyzed cross-coupling reactions in
supercritical carbon dioxide. Chem Commun 1998, (13):1397–1398
60. Shezad N, Oakes RS, Clifford AA et al (1999) Use of fluorinated palladium sources for
efficient Pd-catalysed coupling reactions in supercritical carbon dioxide. Tetrahedron Lett
40(11):2221–2224
61. Early TR, Gordon RS, Carroll MA et al (2001) Palladium-catalysed cross-coupling reactions
in supercritical carbon dioxide. Chem Commun 1(19):1966–1967
62. Kuchurov IV, Vasil’ev AA, Zlotin SG (2010) The Suzuki–Miyaura cross-coupling of bromo-
and chloroarenes with arylboronic acids in supercritical carbon dioxide. Mendeleev Commun
20(3):140–142
63. Fernandes RR, Lasri J, da Silva MFCG et al (2011) Oxadiazoline and ketoimine Palladium
(II) complexes as highly efficient catalysts for Suzuki–Miyaura cross-coupling reactions in
supercritical carbon dioxide. Adv Synth Catal 353(7):1153–1160
64. Grubbs RH, Miller SJ, Fu GC (1995) Ring-closing metathesis and related processes in
organic synthesis. Acc Chem Res 28(11):446–452
65. Schrock RR, Murdzek JS, Bazan GC et al (1990) Synthesis of molybdenum imido alkylidene
complexes and some reactions involving acyclic olefins. J Am Chem Soc 112(10):3875–3886
66. Fürstner A, Koch D, Langemann K et al (1997) Olefin metathesis in compressed carbon
dioxide. Angew Chem Int Ed 36(22):2466–2469
67. Theruvathu JA, Aravindakumar CT, Flyunt R et al (2001) Fenton chemistry of
1,3-dimethyluracil. J Am Chem Soc 123(37):9007–9014
68. Mikami K, Matsukawa S, Kayaki Y et al (2000) Asymmetric Mukaiyama aldol reaction of a
ketene silyl acetal of thioester catalyzed by a binaphthol–titanium complex in supercritical
fluoroform. Tetrahedron Lett 41(12):1931–1934
69. Komoto I, Kobayashi S (2001) Lewis acid catalysis in a supercritical carbon dioxide (scCO2)-
poly(ethylene glycol) derivatives (PEGs) system: remarkable effect of PEGS as additives on
13 Homogeneous Catalysis Promoted by Carbon Dioxide 367

reactivity of Ln(OTf)-catalyzed Mannich and Aldol reactions in scCO2. Chem Commun


2001, (18):1842–1843
70. Komoto I, Kobayashi S (2004) Lewis acid catalysis in supercritical carbon dioxide. Use of
poly(ethylene glycol) derivatives and perfluoroalkylbenzenes as surfactant molecules which
enable efficient catalysis in scCO2. J Org Chem 69(3):680–688
71. Kayaki Y, Noguchi Y, Iwasa S et al (1999) An efficient carbonylation of aryl halides
catalysed by palladium complexes with phosphite ligands in supercritical carbon dioxide.
Chem Commun 1999, (13):1235–1236
72. Pauson PL (1985) The khand reaction: a convenient and general route to a wide range of
cyclopentenone derivatives. Tetrahedron 41(24):5855–5860
73. Schore NE (1988) Transition metal-mediated cycloaddition reactions of alkynes in organic
synthesis. Chem Rev 88(7):1081–1119
74. Jeong N, Hwang SH, Lee YW et al (1997) Catalytic Pauson–Khand reaction in super critical
fluids. J Am Chem Soc 119(43):10549–10550
75. Li G, Gao J, Wei H-X et al (2000) New CC bond formation via nonstoichiometric Titanium
(IV) halide mediated vicinal difunctionalization of α, β-unsaturated acyclic ketones. Org Lett
2(5):617–620
76. Shi M, Xu Y-M (2003) An unexpected highly stereoselective double aza-Baylis–Hillman
reaction of sulfonated imines with phenyl vinyl ketone. J Org Chem 68(12):4784–4790
77. Rose PM, Clifford AA, Rayner CM (2002) The Baylis-Hillman reaction in supercritical
carbon dioxide: enhanced reaction rates, unprecedented ether formation, and a novel phase-
dependent 3-component coupling. Chem Commun 2002, (9):968–969
78. Cheng J-S, Jiang H-F (2004) Palladium-catalyzed regioselective cyclotrimerization of
acetylenes in supercritical carbon dioxide. Eur J Org Chem 2004(3):643–646
79. Montilla F, Avilés T, Casimiro T et al (2001) CpCo(CO)2-catalysed cyclotrimerisation of
alkynes in supercritical carbon dioxide. J Organomet Chem 632(1–2):113–118
80. Musie G, Wei M, Subramaniam B et al (2001) Catalytic oxidations in carbon dioxide-based
reaction media, including novel CO2-expanded phases. Coord Chem Rev 219–221:789–820
81. Maayan G, Ganchegui B, Leitner W et al (2006) Selective aerobic oxidation in supercritical
carbon dioxide catalyzed by the H5PV2Mo10O40 polyoxometalate. Chem Commun 2006
(21):2230–2232
82. Leitner W (1999) Reactions in supercritical carbon dioxide. Top Curr Chem 206:107–132
83. Chang Y, Jiang T, Han B et al (2003) Aerobic oxidation of cyclohexanol to cyclohexanone in
compressed CO2 and liquid solvents. Ind Eng Chem Res 42(25):6384–6388
84. Herbert M, Montilla F, Galindo A (2010) Supercritical carbon dioxide, a new medium for
aerobic alcohol oxidations catalysed by copper-TEMPO. Dalton Trans 39(3):900–907
85. Ansari MB, Park S-E (2012) Carbon dioxide utilization as a soft oxidant and promoter in
catalysis. Energy Environ Sci 5(11):9419–9437
86. Gu L, Zhang Y (2010) Unexpected CO2 splitting reactions to form CO with N-heterocyclic
carbenes as organocatalysts and aromatic aldehydes as oxygen acceptors. J Am Chem Soc
132(3):914–915
87. Nair V, Varghese V, Paul RR et al (2010) NHC catalyzed transformation of aromatic
aldehydes to acids by carbon dioxide: an unexpected reaction. Org Lett 12(11):2653–2655
88. Chiang P-C, Bode JW (2011) On the role of CO2 in NHC-catalyzed oxidation of aldehydes.
Org Lett 13(9):2422–2425
89. Ren X, Yuan Y, Ju Y et al (2012) Oxidation ability of CO2 for the transformation of cinnamic
aldehydes to acids catalyzed by N-heterocyclic carbene: combining computational and
experimental studies. ChemCatChem 4(12):1943–1951
90. Olsen MHN, Salomão GC, Drago V et al (2005) Oxidation of cyclohexane in supercritical
carbon dioxide catalyzed by iron tetraphenylporphyrin. J Supercrit Fluids 34(2):119–124
91. Kokubo Y, Wu X-W, Oshima Y et al (2004) Aerobic oxidation of cyclohexene catalyzed by
Fe(III)(5,10,15,20-tetrakis(pentafluorophenyl)porphyrin)Cl in supercritical CO2. J Supercrit
Fluids 30(2):225–235
368 R. Ma et al.

92. Nolen SA, Lu J, Brown JS et al (2002) Olefin epoxidations using supercritical carbon dioxide
and hydrogen peroxide without added metallic catalysts or peroxy acids. Ind Eng Chem Res
41(3):316–323
93. Jiang H, Jia L, Li J (2000) Wacker reaction in supercritical carbon dioxide. Green Chem
2(4):161–164
94. Jiang H-F, Shen Y-X, Wang Z-Y (2008) Palladium-catalyzed aerobic oxidation of terminal
olefins with electron-withdrawing groups in scCO2. Tetrahedron 64(3):508–514
95. ten Brink GJ, Arends IWCE, Sheldon RA (2004) The Baeyer–Villiger reaction: new devel-
opments toward greener procedures. Chem Rev 104(9):4105–4124
96. Bolm C, Palazzi C, Francio G et al (2002) Baeyer–Villiger oxidation in compressed CO2.
Chem Commun 2002, (15):1588–1589
97. DeSimone JM, Guan Z, Elsbernd CS (1992) Synthesis of fluoropolymers in supercritical
carbon dioxide. Science 257:945–947
98. Clark MR, DeSimone JM (1995) Cationic polymerization of vinyl and cyclic ethers in
supercritical and liquid carbon dioxide. Macromolecules 28(8):3002–3004
99. Albertsson A-C, Varma IK (2003) Recent developments in ring opening polymerization of
lactones for biomedical applications. Biomacromolecules 4(6):1466–1486
100. Bratton D, Brown M, Howdle SM (2005) Tin(II) ethyl hexanoate catalyzed precipitation
polymerization of ε-caprolactone in supercritical carbon dioxide. Macromolecules
38(4):1190–1195
101. Hamilton JG, Rooney JJ, DeSimone JM et al (1998) Stereochemistry of ring-opened metath-
esis polymers prepared in liquid CO2 at high pressure using Ru(H2O)6(Tos)2 as catalyst.
Macromolecules 31(13):4387–4389
102. Hu X, Blanda MT, Venumbaka SR et al (2005) Ring-opening metathesis polymerization
(ROMP) of norbornene in supercritical carbon dioxide using well-defined metal carbene
catalysts. Polym Advan Technol 16(2–3):146–149
103. Hori H, Six C, Leitner W (1999) Rhodium-catalyzed phenylacetylene polymerization in
compressed carbon dioxide. Macromolecules 32(10):3178–3182
104. Gimenez-Pedros M, Tortosa-Estorach C, Bastero A et al (2006) Alternating CO/tert-
butylstyrene copolymerisation using soluble cationic palladium complexes in supercritical
carbon dioxide. Green Chem 8(10):875–877
Chapter 14
Multiphase Catalytic Reactions in/Under
Dense-Phase Carbon Dioxide: Utilization
of Carbon Dioxide as a Reaction Promoter

Hiroshi Yoshida, Shin-ichiro Fujita, Masahiko Arai,


and Bhalchandra M. Bhanage

14.1 Introduction

Carbon dioxide serves as a building block for organic compounds, and various
interesting examples are shown in the other chapters of this book. Carbon dioxide
may be transformed into value-added organic compounds through chemical, pho-
tochemical, electrochemical, and biochemical reactions. In addition to the role of
CO2 as a useful reactant, CO2 acts as a promoter/modifier in organic reactions in
liquid phases although it is not involved as a reactant. In order for CO2 to show such
actions, pressurization is necessary to dissolve CO2 in a certain quantity in an
organic phase. When an organic liquid (substrate, solvent) is pressurized by CO2,
its volume may increase to some extent depending on the nature of the liquid.
In some cases, a considerable amount of CO2 is dissolved, and this expands the
volume of the organic liquid phase significantly. This phase is often called as CO2-
dissolved expanded liquid phase (CXL) [1–3]. The dissolution of CO2 changes the
solvent properties of a liquid phase from pure organic liquid state to high-density
CO2 state, depending on its pressure applied. The liquid-phase dissolving CO2 is
beneficial to dissolution of gaseous reactants such as oxygen, hydrogen, and carbon
monoxide. Thus, the pressurization with CO2 would accelerate the liquid-phase
chemical reactions involving gaseous reactants, including oxidation, hydrofor-
mylation, and hydrogenation. That is, CO2 can act as an accelerator in organic
synthetic reactions on simple pressurization of the reaction mixtures with CO2

H. Yoshida • S.-i. Fujita • M. Arai (*)


Division of Chemical Process Engineering, Faculty of Engineering, Hokkaido University,
Sapporo 060-8628, Japan
e-mail: marai@eng.hokudai.ac.jp
B.M. Bhanage
Department of Chemistry, Institute of Chemical Technology, Mumbai 400019, India

B.M. Bhanage and M. Arai (eds.), Transformation and Utilization of Carbon Dioxide, 369
Green Chemistry and Sustainable Technology, DOI 10.1007/978-3-642-44988-8_14,
© Springer-Verlag Berlin Heidelberg 2014
370 H. Yoshida et al.

H2 CO O2
• Assists the dissolution of
CO2 gaseous reactants

• Interacts with functional


CO2 groups of substrates and/or
intermediates and modify
CO2 CO2 their reactivity

CO2

H CHO H CH2OH
H2

R H R H

CO2

Fig. 14.1 Multiphase catalytic reactions in the presence of pressurized CO2. The liquid phase
includes substrate, solvent, and catalyst (homogeneous or heterogeneous). The dissolution of CO2
may improve the dissolution of gaseous reactants and modify the reactivity of certain functional
groups of organic substrates. An example given is the selective hydrogenation of carbonyl group,
than carbon-carbon double bond, of α,β-unsaturated aldehyde to unsaturated alcohol

(Fig. 14.1). Under certain conditions, however, the dissolved CO2 molecules act
only as a diluent and have a negative effect of retarding the reactions. Whether the
promotional effect of CO2 pressurization will appear or not depends on a relative
balance between those positive and negative factors.
In addition to the above-mentioned macro (physical) functions, CO2 has inter-
esting and significant micro (chemical) functions in organic synthetic reactions
through direct interactions with reacting species in the liquid phases. The dissolved
CO2 molecules interact with functional groups of substrates and/or reaction inter-
mediates, as proved by in situ high-pressure Fourier transform infrared spectros-
copy [4]. The direct interaction with CO2 changes the reactivity of certain
functional groups of substrates and/or reaction intermediates, and, as a result, this
may improve the selectivity to the production of desired products. That is, CO2 acts
as a reaction modifier, which is expected for liquid-phase synthetic reactions
including no gaseous reactants (e.g., Heck coupling) as well as those including
gaseous reactants. This chapter reviews several interesting examples of micro
(chemical) effects of CO2 in organic synthetic reactions and possible explanations
for these effects observed. The reactions dealt with this chapter include oxidation,
hydroformylation, hydrogenation, acid-catalyzed reactions, Heck reactions, and
Diels–Alder reactions.
14 Multiphase Catalytic Reactions in/Under Dense-Phase Carbon Dioxide. . . 371

14.2 Multiphase Catalytic Reactions in the Presence


of Dense-Phase CO2

14.2.1 Oxidation

Oxidation is one of the widely investigated reactions in both academia and industry,
and O2 and H2O2 are generally used as oxidants. The use of dense-phase CO2
prevents oxidation reactions from combustion and explosion, and mass transfer
limitation can be improved in multiphase catalytic oxidations using gaseous
oxidants. Therefore, several researchers investigated the utilization of CO2 in
oxidation reactions.
Yoo et al. reported the chemical functions of CO2 as an active component for the
liquid-phase O2 oxidation of o-, m-, and p-xylenes at 468 K with Co/Mn/HBr
catalyst, which is called “MC-type O2 oxidation” [5]. They showed that the amount
of O2 consumed in O2 oxidation of m-xylene was enhanced by the copresence of
CO2, and the purity of isophthalic acid (IPA) in the recovered solid product was also
increased. Moreover, the promoting effects of CO2 increased with increasing
concentration of CO2, indicating that both catalytic activity and selectivity to IPA
were influenced by the presence of CO2 in the reaction media. The O2 oxidation of
p-toluic acid, which was produced as an intermediate from p-xylene, was also
examined, and they concluded that there were three mechanistic interpretations of
the effects of CO2. Firstly, copresence of CO2 enhanced the solubility of the
substrates as well as the intermediates; secondly, decarboxylation of the interme-
diates can be suppressed by the presence of CO2. Thirdly, a synergistic interaction
of CO2 and O2 molecules may create a cyclic peroxocarbonate species on the metal
center of the MC-type catalyst, and these species play a key role in initiating the
free radical propagation via the hydrogen abstraction process and accelerating
the oxidation. The possible pathways of formation of peroxocarbonate species are
shown in Scheme 14.1.
The metal-O2 complex is initially formed on either Co or Mn center, and the CO2
molecule is then inserted into the O-O bond to produce an active species (Type I).

Scheme 14.1 Possible pathways of formation of peroxocarbonate species from O2 and CO2 with
metal center of MC-type catalyst
372 H. Yoshida et al.

Scheme 14.2 Reaction pathways of O2 oxidation of xylenes over Co/Mn/HBr catalyst in the
presence of CO2

Scheme 14.3 The


formation of
peroxycarbonic acid from
CO2 and H2O2

Another type of peroxocarbonate species can also be formed by a different mode of


interaction (Type II). Aresta et al. reported that the oxygen atom directly bonded to
the metal center in these species can transfer to the oxophile substrates [6],
indicating that the peroxocarbonate species formed from O2 and CO2 accelerated
the oxidation of methyl group in aromatic compounds to corresponding aldehydes
as shown in Scheme 14.2.
Eckert et al. have investigated the olefin epoxidations using H2O2 as oxidant in
the presence of dense-phase CO2 with no metallic catalyst [7] and mentioned
the formation of peroxycarbonic acid from CO2 and H2O2, which facilitates the
olefin epoxidation (Scheme 14.3). The mass transfer limitation is one of the crucial
problems in this reaction because the epoxidation occurs in aqueous phase. Dense-
phase CO2 contributes to improve the mass transfer. Subramaniam et al. indicated
that the peroxycarbonate is the most likely reactive intermediate in the olefin
epoxidation [8]. In addition, they note that hydrogen bonding and polarity of
CXLs are key factors in the epoxidation, which can be controlled by CO2 pressur-
ization and the kind of organic solvent used.

14.2.2 Hydroformylation

Hydroformylation of olefin with H2 and CO is an industrially important reaction,


and homogeneous Rh or Co complex catalysts are generally used. The cost of Rh
complex catalyst is relatively expensive than Co complex catalyst, but the reaction
can occur with high activity and selectivity under milder conditions. The utilization
of dense-phase CO2 in hydroformylation has been investigated by several
researchers [9–11]. It is indicated that the increase of the solubility of syngas into
liquid phase by using CXLs as reaction media gives the most significant effect in
the hydroformylation reactions.
Tominaga et al. have first shown that the hydroformylation can indeed proceed
with a mixture of H2 and CO2 instead of CO [12, 13]. CO can be formed from H2
14 Multiphase Catalytic Reactions in/Under Dense-Phase Carbon Dioxide. . . 373

Scheme 14.4 One-pot


hydroformylation of
cyclohexene using H2
and CO2

and CO2 through the reverse water gas shift reaction, and it can be directly used for
hydroformylation with H2, indicating that one-pot hydroformylation from H2 and
CO2 can be achieved. Fujita et al. investigated the effect of H2 and CO2 pressures on
the hydroformylation of cyclohexene with a catalyst system of Ru3(CO)12 and LiCl
in N-methyl-2-pyrrolidone [14]. The hydroformylation with CO2 and H2 has a
two-step process, as shown in Scheme 14.4, and the aldehyde produced is further
reduced to the corresponding alcohol. In their work, gaseous CO was detected
during the reaction, and its amount was comparable to those of aldehyde and
alcohol formed. The step 1 proceeds faster compared with the step 2, which
means that the hydroformylation reaction is the rate-determining step. The forma-
tion of CO was accelerated with increasing total pressure of H2 and CO2, and the
yield of cyclohexanecarboxaldehyde was also increased. However, both the hydro-
genation of cyclohexanecarboxaldehyde to cyclohexane-methanol and that of
cyclohexene to cyclohexane were suppressed with increasing CO2 pressure. The
chemical functions of CO2 as a promoter as well as a reactant were examined by in
situ high-pressure FTIR measurements, and molecular interactions of CO2 with
cyclohexene and intermediates were not observed. They also examined the struc-
ture of Ru complexes at different pressures of H2 and CO2, and the results indicated
the presence of [H3Ru4(CO)12] 1 and [HRu3(CO)11] 1 under the reaction condi-
tions used. These species may be active for the reduction of CO2 to CO and the
following hydroformylation reaction, respectively.

14.2.3 Hydrogenation

Liquid-phase hydrogenation is one of industrially important reactions, and different


homogeneous and heterogeneous catalysts are used [15, 16]. Heterogeneous cata-
lysts are more desirable than homogeneous ones when we consider the post-
reaction treatments of catalyst separation and recycling. The reaction mixtures of
our interest here are a multiphase system containing a liquid phase (substrate,
solvent) and/or a gas phase (H2, CO2) with a solid phase (heterogeneous catalyst).
The pressurization of organic liquid phases with CO2 may have an interesting
impact on the reaction rate and/or the product selectivity for the hydrogenation
reactions in these phases.
374 H. Yoshida et al.

Unsaturated alcohol
Step1 H CH2OH
H2 H2
H CHO R H H CH2OH

R H H2 H CHO R H
H2
α,β-Unsaturated Saturated alcohol
Step2 H
aldehyde R
Saturated aldehyde

Scheme 14.5 Reaction pathways in hydrogenation of α,β-unsaturated aldehyde

The acceleration of hydrogenation reactions in the presence of dense-phase CO2


was reported in the literature [17–20]. For example, Chouchi et al. compared the
rates of hydrogenation of α-pinene with a Pd/C catalyst under different conditions
at 323 K. They showed that the rate of hydrogenation in a gas–liquid system under
CXL conditions was even faster compared to that in a homogeneous scCO2
system [20].
In selective hydrogenation of α,β-unsaturated aldehydes with supported Pt
catalysts, the C═C group is more easily hydrogenated rather than the C═O group,
and so it is difficult to produce the corresponding unsaturated alcohols in high
selectivity [21] (Scheme 14.5). The high selectivity to unsaturated alcohols may be
achieved by modifying the properties of supported Pt particles with Sn and other
additives. Bhanage et al. revealed that the presence of dense-phase CO2 was
effective for improving the hydrogenation of the carbonyl bond of cinnamaldehyde,
α-methyl-trans-cinnamaldehyde, and crotonaldehyde with a Pt/Al2O3 catalyst at
323 K [22]. For crotonaldehyde, for example, the selectivity to crotyl alcohol was
46 % in neat ethanol at a H2 pressure of 4 MPa, while it was improved to >70 % in
the presence of 14 MPa CO2. For cinnamaldehyde, the selectivity to cinnamyl
alcohol was enhanced from 78 % in ethanol to 93 % at a CO2 pressure of 14 MPa; in
addition, the rate of hydrogenation was also improved with increasing CO2 pres-
sure. Zhao et al. measured in situ FTIR spectra of cinnamaldehyde and benzalde-
hyde dissolved in pressurized CO2 gas phase at 323 K [23, 24]. They observed a
significant red shift of the frequency of the stretching vibration of the C═O bond
with increasing CO2 pressure for cinnamaldehyde but a marginal red shift for
benzaldehyde. The frequency of the vibration of the C═C bond was found to
change little with the pressure. Therefore, they assumed that CO2 molecules were
able to interact with the carbonyl group, decrease the bond strength, and then
increase its reactivity to hydrogenation. Those molecular interactions between
CO2 species and the carbonyl group were studied and confirmed by theoretical
model calculations by Zhao et al. [25]. The function of CO2 as an accelerator/
modifier was also observed for the same hydrogenation reactions of α,β-unsaturated
aldehydes using homogeneous metal complex catalysts [26–28].
The dense-phase CO2 also has an impact on liquid-phase hydrogenation of
aromatic nitro compounds, which is important in industry for the manufacture of
aniline compounds [29], and several metals are used as active catalysts [30–34].
14 Multiphase Catalytic Reactions in/Under Dense-Phase Carbon Dioxide. . . 375

NO2 NO NHOH NH2


H2 H2 H2

Step 1 Step 2 Step 3


Nitrobenzene Nitrosobenzene N-phenylhydroxylamine Aniline

Scheme 14.6 Hydrogenation of nitrobenzene toward aniline

In this hydrogenation (Scheme 14.6), the production of such intermediates as


nitrosobenzene and N-phenylhydroxylamine and such coupling by-products as
azoxybenzene, azobenzene, and hydrazobenzene should be avoided, and a high
selectivity to aniline should be achieved at a high conversion.
Meng et al. studied the selective hydrogenation of nitrobenzene to aniline with
Al2O3-supported Ni catalysts at 323 K and at a H2 pressure of 4 MPa [35]. The
reaction runs were conducted in the presence of CO2 at different pressures and no
additional organic solvent. Under the conditions used, the rate of hydrogenation
was observed to increase with increasing CO2 pressure up to about 12 MPa, below
which the reaction mixture was a gas–liquid–solid three-phase system and the
reaction occurred mainly in the liquid phase. The dissolution of a gaseous reactant
of H2 should be improved by CO2 pressurization, and this effect may be superior to
the dilution of the reacting species in the liquid phase. The reaction mixture
changed to a gas-solid two-phase system at higher CO2 pressure, and the dilution
became more significant, resulting in a decrease in the rate of reaction. It is
noteworthy that the selectivity to the desired product of aniline is almost 100 %
at any conversion level, and this complete selective hydrogenation can be achieved
at CO2 pressures of 6–18 MPa, that is, irrespective of the phase behavior of the
reaction mixture. Meng et al. measured in situ FTIR spectra of nitrobenzene,
nitrosobenzene, and N-phenylhydroxylamine dissolved in dense-phase CO2 at
different pressures. They observed that the frequency of stretching vibration of
N–O bond of the nitro group was blue-shifted on CO2 pressurization and believed
that the reactivity of the nitro group was reduced in the presence of dense-phase
CO2. They assumed that for nitrosobenzene and N-phenylhydroxylamine, in con-
trast, their reactivity was promoted by the action of CO2 species in the liquid phase.
Thus, CO2 is believed to slow down the rate of step 1 but accelerate the rates of
steps 2 and 3. As a result, aniline can be produced in 100 % selectivity with no
undesired intermediates and coupling products. The dilution of intermediates of
nitrosobenzene and N-phenylhydroxylamine by the dissolution of CO2 and H2
would contribute to the suppression of their coupling reactions.
Meng et al. made a similar study with chloronitrobenzene substrates [36]. They
demonstrated that the pressurization with CO2 is also effective for the 100 %
selective hydrogenation to the desired chloroaniline products at any conversion
level. No dechlorination was observed to occur in this multiphase reaction system.
Zhao et al. investigated the effects of metal particle size in hydrogenation of
nitrobenzene with several supported noble metal catalysts in the presence of
376 H. Yoshida et al.

Dibenzylamine

N
H

Step 4
−NH3

N NH NH2

H2 H2

Step 1 Step 2
Benzonitrile Benzylideneimine CO2 Benzylamine
Step 3
+ O −

NH3 N O
H

Carbamate salt

Scheme 14.7 Reaction pathways in hydrogenation of benzonitrile

dense-phase CO2 [37, 38]. They showed that the turnover frequency decreased with
the degree of metal dispersion in similar fashions for the reactions in ethanol and in
compressed CO2.
An interesting result was further reported by Meng et al. that the positive
effect of CO2 pressurization on the selective hydrogenation of nitrobenzene to
aniline appeared at a low CO2 pressure of about 1 MPa when an additional phase
of water was also added to the reaction system [39]. Lin et al. reported the
effectiveness of compressed CO2/water system for ring hydrogenation of aromatic
alcohols and aldehydes [40]. The function of water added is not clear at present. The
multiphase organic reaction systems including dense-phase CO2 and water are
interesting and worth investigating in detail. A recent review article demonstrates
the significance of on-water organic reactions, in which organic-water interface is
important [41]. The combination of dense-phase CO2 and water could give additive
or synergistic effects on the reaction rate and/or the product selectivity in organic
synthetic reactions.
Yoshida et al. reported different CO2 actions for hydrogenation of benzonitrile
with a Pd/Al2O3 catalyst in multiphase media including dense-phase CO2 and water
at 323 K [42]. The selective hydrogenation of benzonitrile to the primary amine
may be achieved in a selectivity >95 % in a medium including both CO2 and water.
In this system, the primary amine formed may easily transform with CO2 to a water-
soluble carbamate salt (Scheme 14.7), and this hinders its further hydrogenation to
the secondary amine in the organic phase, resulting in the high selectivity to
benzylamine. The carbamate salt is not stable, and so the benzylamine can be
14 Multiphase Catalytic Reactions in/Under Dense-Phase Carbon Dioxide. . . 377

H2 H2

Step 1 Step 2
Phenylacetylene Styrene Ethylbenzene

Scheme 14.8 Hydrogenation of phenylacetylene to styrene and ethylbenzene

Scheme 14.9 Vinylaniline


Hydrogenation of the nitro NH2
and vinyl groups of
nitrostyrene
NO2 H2 H2 NH2

H2 NO2
H2
Nitrostyrene Ethylaniline

Ethylnitrobenzene

regenerated by heating [43]. When water is absent, the carbamate salt deposits on
the surface of catalyst and causes its deactivation. The primary amine is soluble
somewhat in water, but the production of the secondary amine cannot be avoided in
the absence of CO2.
The action of CO2 as a protecting agent was earlier reported for hydrogenation
[44] and other synthetic reactions [45, 46] using homogeneous catalysts. Eckert
et al. studied the selective hydrogenation of benzonitrile with NiCl2/NaBH4 in
ethanol at 303 K [44]. The yield of the primary amine was markedly promoted by
simple pressurization of the organic phase with CO2. This is ascribable to the
formation of carbamate salt that prevents the side reaction of the primary amine
with imine intermediate giving the secondary amine.
Recently Wei et al. investigated the partial hydrogenation of phenylacetylene
with several monometallic and bimetallic Pd-based catalysts in hexane at 313 K
[47] (Scheme 14.8). The pressurization with CO2 was shown to improve the
selectivity to styrene; at a total pressure of 3.4 MPa including 20 % CO2, styrene
was produced in a selectivity of 92 % at 100 % conversion in 0.5 h with a Pd-Ag
bimetallic catalyst. Further increase of CO2 partial pressure to 40 % increased the
selectivity to 98 % but decreased the conversion to 74 % under the conditions used.
Namely, the further hydrogenation of an alkene of styrene to an alkane of ethyl-
benzene can be suppressed although the function of CO2 molecules is still unclear.
The extent of impact of CO2 pressurization should be different from one organic
liquid to another depending on their solvent properties; above-mentioned positive
effects would be expected for some organic liquids but not for others. The disso-
lution of a certain amount of CO2 into a liquid is required for CO2 to have positive
(or negative) effects in the liquid phase. Yoshida et al. studied the effects of CO2 for
a test reaction of nitrostyrene hydrogenation (Scheme 14.9) with a Pt/TiO2 catalyst
378 H. Yoshida et al.

using two organic solvents of toluene and ethanol at 323 K [48]. The two solvents
indicated similar degrees of volume expansion at the same CO2 pressure, as
confirmed by phase behavior observations at the reaction temperature. In neat
toluene, neat ethanol, and ethanol pressurized by CO2, the nitro group of
nitrostyrene was hydrogenated to vinylaniline in a high selectivity. In contrast,
the vinyl group was selectively hydrogenated to ethylnitrobenzene in toluene
pressurized by CO2. These results may be explained by relative interactions
among the two functional groups, the solvent molecules, and the CO2 species in
the liquid phase. The interactions of the vinyl group with CO2 are weak, but those of
the nitro group with CO2 are significant, which decrease its reactivity to hydroge-
nation. The interactions of the nitro group with the dissolved CO2 can appear in the
case of toluene, in which the nitro group is likely to be surrounded by CO2 species
rather than toluene molecules because of no interaction between the nitro group and
toluene molecules. So, the impact of CO2 pressurization appears even at low
pressures, decreasing the reactivity of the nitro group and then relatively increasing
the hydrogenation of the vinyl group. In the case of ethanol, however, the inter-
actions of the nitro group with the solvent molecules are strong as compared to
those with CO2, and, hence, the impact of CO2 pressurization does not appear even
at a high pressure of 10 MPa. These explanations are supported by molecular
dynamics simulations, which determine the local composition around the nitro
group of nitrostyrene in either toluene or ethanol pressurized by CO2. It is thus
important to select a suitable solvent when we expect positive effects of CO2
pressurization on liquid-phase hydrogenation and other organic reactions. Solvent
properties of substrates, solvents, and CO2 are important to consider [49].
Those results demonstrate the importance of interactions of CO2 molecules with
substrates and intermediates. Akiyama et al. studied these molecular interactions of
CO2 with various organic compounds of ketones, esters, and amides by using in
situ high-pressure FTIR [4]. The results would be helpful in discussing the impact
of CO2 on reactions with those substrates. In the hydrogenation reactions in the
presence of dense-phase CO2 over supported metal catalysts, CO is formed from
CO2 and H2 and adsorbed on the surface of supported metal particles, and this
would cause the deactivation of catalysts [50–54]. The occurrence of catalyst
deactivation may depend on competitive adsorption of CO, substrates, and/or
intermediates. Dong et al. studied the hydrogenation of the aromatic rings of
polystyrene in a CXL of decahydronaphthalene pressurized by CO2 using 5 %
Pd/SiO2 and 5 % Pd/Al2O3 catalysts at 423 K [55]. They observed the catalyst
deactivation in a relatively early stage of reaction due to the formation and
adsorption of CO formed from CO2 and H2. They detected CO at the end of
hydrogenation reaction. It is interesting to note that the catalyst deactivation can
be avoided by physically mixing an additional methanation catalyst (5 % Ru/Al2O3,
65 % Ni/SiO2-Al2O3) and a hydrogenation catalyst into a single particle. The
methanation catalyst could convert CO into CH4 and regenerate the surface of the
hydrogenation catalysts. The use of the methanation catalysts does not influence
the desired reaction of the aromatic ring hydrogenation of polystyrene. Using in situ
high-pressure FTIR, Yoshida et al. studied the formation and adsorption of CO
14 Multiphase Catalytic Reactions in/Under Dense-Phase Carbon Dioxide. . . 379

Scheme 14.10 In situ


formation of alkylcarbonic
acid from CO2 and alcohol

Scheme 14.11 Formation of dimethyl acetal of cyclohexanone catalyzed by in situ generated acid
in CO2-expanded methanol

from H2 and CO2 over Al2O3-supported Pt, Pd, Rh, and Ru catalysts at 323 K
[56]. They collected FTIR spectra of adsorbed CO at different CO2 pressures and in
the presence and absence of water (H2O, D2O) for heat-treated and untreated noble
metal catalysts. They discussed the types of CO species adsorbed on the metal
particles depending on their surface structure and the influence of water vapor on
the CO adsorption behavior. Water molecules may be adsorbed more preferentially
on atomically rough surfaces rather than CO species. Our attention is herein given
to interactions of CO2 with organic substrates and intermediates. The CO2 may
affect the properties of supported metal particles through the formation and adsorp-
tion of CO in hydrogenation reactions, as mentioned above. Note, further, that CO2
molecules could have a direct impact on supported metal particles, which was
speculated from the optical absorption measurements of supported Au particles in
the compressed CO2 gas phase at different pressures [57]. Liu et al. recently
reported that the properties and activity of a TiO2 catalyst were modified by the
accumulation of CO2 species on its surface assisted by the action of a surfactant
[58]. The photocatalytic activity of TiO2 for the production of H2 through water
splitting was improved by about one order of magnitude.

14.2.4 Acid-Catalyzed Reactions

Acid-catalyzed reactions in CXLs using in situ generated catalysts from the reac-
tion of CO2 with alcohol or water were pioneered by the joint group of Eckert and
Liotta [59–63]. Alkylcarbonic acids can be formed in alcohols with compressed
CO2 and used as acid catalysts (Scheme 14.10). This reaction system is desirable
because the decomposition of the acid can automatically occur by simple depres-
surization after the reactions without any post-reaction neutralization and salt
disposal.
They found that the formation of dimethyl acetal of cyclohexanone in methanol
(Scheme 14.11) was up to 130 times faster by CO2 pressurization. The greatest
380 H. Yoshida et al.

Scheme 14.12 Hydrolysis of β-pinene in the CO2-expanded MeOH/H2O mixture

Scheme 14.13 In situ acid catalysis for single-pot synthesis of methyl yellow and iodobenzene in
CO2-expanded methanol

enhancement was found at a pressure of 22 bar; presumably higher pressures are


undesirable due to decreasing polarity of the expanded solvent inhibiting the
dissociation of the acid.
The application of in situ generated alkylcarbonic acid as a catalyst for the
hydrolysis of β-pinene to terpineol and other alcohols was also investigated
[61]. Good selectivity for alcohols rather than hydrocarbons was obtained in
CXLs and hot water (Scheme 14.12). Eckert et al. used several kinds of CXLs
with different compositions of methanol, water, and CO2 for the reactions.
A selectivity of 84.0 % for the hydrolysis product at the highest conversion
(70.9 %) was achieved in the CO2-expanded 1:1 methanol/water mixture (348 K,
3 mol% CO2). The reaction did not proceed when the CO2 was absent.
Eckert et al. demonstrated the catalytic performance of in situ acid catalysis for
the diazotization of aniline, which is either coupled with N,N-dimethyl aniline to
form methyl yellow or reacted with iodide to form iodobenzene (Scheme 14.13)
[63]. A 97 % conversion of aniline with a 72 % yield of methyl yellow was obtained
at 278 K in 72 h with an excess nitrile salt and a high CO2 loading. This reaction
did not occur in the absence of CO2. They also used THF as a solvent instead of
methanol, but the yield was only 0.3 %, which may come from the lack of
alkylcarbonic acid and the formation of carbamates from aniline and CO2. These
results strongly suggest that a suitable acid source like alcohol must be present for
utilization of dense-phase CO2 in the acid-catalyzed reactions.
14 Multiphase Catalytic Reactions in/Under Dense-Phase Carbon Dioxide. . . 381

Scheme 14.14 Heck


reactions of halogenated
benzene derivatives with
methyl acrylate in CO2-
expanded toluene with
Pd/TPP catalyst

14.2.5 Heck and Diels–Alder Reactions

There are many advantages of using dense-phase CO2 in multiphase catalytic


reactions as mentioned above. Recently, some positive effects in liquid organic
reactions without any gaseous reactants were also found.
ScCO2 has been investigated as a reaction media of Heck coupling reactions
using homogeneous complex catalysts; there are some advantages such as the
extraction of the product, resulting in the controlling of the product distribution.
The modification of metal complex catalysts with fluorinated functional groups is
effective for increasing their solubility into scCO2 phase [64, 65]. Such a catalyst
modification may not always be necessary for multiphase reactions, in which the
solubility of the catalysts in the organic phase is not so a serious problem.
Multiphase Heck reactions in the presence of dense-phase CO2 were investigated
[66]. Fujita et al. reported the effects of CO2 pressurization on liquid-phase Heck
reactions of methyl acrylate with various aryl bromides over homogeneous Pd/TPP
catalyst in toluene in the presence of compressed CO2 (Scheme 14.14) [67]. For the
Heck reaction of iodobenzene and non-activated bromobenzene with methyl acry-
late, the conversion simply decreased with CO2 pressure due to the dilution effect of
CO2 molecules in the liquid organic phase. However, the conversion increased with
CO2 pressure in the range up to 3 MPa for the reaction of 2-bromoacetophenone
with methyl acrylate. This promoting effect was not observed in the case of using 3-
and 4-bromoacetophenone. From the results of high-pressure FTIR measurements
of 2-, 3-, 4-bromoacetophenones in the presence of CO2, they conclude that the
dissolved CO2 molecules interact with the aryl bromide substrates in different
manners depending on their structures. It is suggested that the C-Br bonds of the
three substrates are modified by the presence of CO2, indirectly through interactions
of CO2 with the other moieties or directly with the C-Br moiety. The rate of the
liquid-phase Heck reaction can be enhanced by pressurizing the reaction mixture
with CO2 at a low pressure of 2–4 MPa, although the enhancement effect is limited
to a few bromoarenes of a certain structure.
Fujita et al. studied the impact of CO2 pressurization for the Diels–Alder
reactions of isoprene with methyl acrylate, methyl vinyl ketone, and acrolein in
the CO2-expanded toluene and ethanol at 393 K with a heterogeneous SiO2 · Al2O3
catalyst [67] (Scheme 14.15). The conversion simply decreased with CO2 pressure
in the reactions with three dienophile substrates, which may be caused by the
dilution effect of CO2 molecules in the liquid phases. However, the product
selectivity remarkably changed with CO2 pressure; the ratio of [3] adduct to [2]
382 H. Yoshida et al.

Scheme 14.15 Two


approaches (a) and (b),
yielding [2] adduct and
[3] adduct

adduct could be altered, which changed most extensively for acrolein from 47:53 to
73:27 at pressure up to 16 MPa. The change was less significant for the other
dienophiles, but the ratio did also change from 70:30 to 75:25 and from 63:37 to
73:27 for methyl acrylate and methyl vinyl ketone, respectively. At a high pressure
of 16 MPa, very similar 3 to 2 ratios were seen for the three dienophiles, which are
similar to those with Diels–Alder reactions of cyclopentadiene with methyl vinyl
ketone and methyl acrylate in dense-phase CO2 reported by Tester et al. [68]. They
concluded that the product selectivity was not affected by CO2 pressure, while
Ikushima et al. mentioned about the effect of CO2 pressurization in the reaction of
isoprene and methyl acrylate in dense-phase CO2 [69]. The pressurization and
dissolution of CO2 in the toluene phase will scarcely change its features as a
continuum of solvent for the reactions. CO2 molecules dissolved surround the
substrate molecules inducing interactions between the substrate and CO2 molecules
in the liquid phase. As the pressure is raised, the state of such a clump of substrate-
CO2 molecules in toluene may become similar to that in a dense-phase CO2
medium, resulting in the similar 3 to 2 ratios as obtained.
The molecular interaction of CO2 with three dienophiles in the presence of
dense-phase CO2 was discussed using in situ high-pressure FTIR measurements,
and the results indicated that the interactions of dense-phase CO2 with the carbonyl
group may be stronger for acrolein compared with the other two carbonyl com-
pounds, as estimated from the peak shift of ν(C═O) absorption band with increasing
CO2 pressure. Fujita et al. speculated that CO2 molecules interact with the carbonyl
group of a dienophile, and this makes its C-1 carbon more positive compared to the
C-2 carbon. As a result, the approach b of Scheme 14.15 may be more favorable to
occur than that between the dienophile and isoprene in which C-3 may be more
positive than C-2 owing to the electron-withdrawing nature of the methyl group
attached to C-2. These interactions may cause steric effects on the product selec-
tivity, resulting in the increase of the 3 to 2 ratio with CO2 pressure. Eckert
et al. also investigated the interactions between CO2 and the substrates for the
Diels–Alder reaction between 4-phenyl-1,2,4-triazoline-3,5-dione (PTAD) and
anthracene in CO2-expanded acetonitrile [70] (Scheme 14.16). From the reaction
results, they speculate that CO2 dissolved in the liquid organic phase acts as a Lewis
acid, and the interaction between Lewis acid and carbonyl oxygen of the PTAD
would destabilize the azo bond, increasing the reactivity of the dienophile toward
the diene (Scheme 14.17).
14 Multiphase Catalytic Reactions in/Under Dense-Phase Carbon Dioxide. . . 383

Scheme 14.16 Diels–Alder reaction of anthracene and PTAD

Scheme 14.17 Lewis


acid–base interaction
between CO2 and PTAD
molecules

Scheme 14.18 The catalytic carbonylative cycloaddition between norbornene 1 and allyl bromide
2 in CO2-expanded acetone

14.2.6 Other Reactions

Recently, many researchers are making efforts to extend the application of dense-
phase CO2 for new chemical reactions including multiphase reactions and liquid-
phase organic reactions.
Moral et al. reported excellent chemoselectivity in the Ni-catalyzed cyclocarbo-
nylation reaction with both alkynes and strained alkenes by tuning the operational
parameters, using CO2-expanded acetone as a reaction medium instead of liquid
acetone [71]. The carbonylative cycloaddition between norbornene 1 and allyl
bromide 2 with CO over Ni/Fe catalyst at 298 K in liquid or CO2-expanded acetone
was investigated (Scheme 14.18). The norbornene and allyl bromide tested were
highly soluble in both pure CO2 and CO2-expanded acetone, but the catalyst was
only soluble in CO2-expanded acetone under certain conditions. Thus, the reaction
was feasible under homogeneous conditions in the latter medium. Cyclopentanone
3 was obtained as a main product in a selectivity of 77 % in the absence of CO2, and
the selectivity to 3 can slightly be improved to 80 % by CO2 pressurization.
Interestingly, they added a certain amount of H2O into the reaction system, and
the selectivity to 3 was enhanced to 98 % with no retardation of the reaction. They
speculated the acidification of reaction media with CO2 and H2O, which decreased
the OH concentration, resulting in preventing hydroxyl attack at the acyl metal
intermediate and interrupting the cyclization.
384 H. Yoshida et al.

Scheme 14.19 Menschutkin reaction of TBA and MNBS

Ford et al. focused on the controlling of local polarity in CXLs by changing CO2
pressure. They examined several probes as indicators of either local or bulk polarity
in CO2-expanded acetonitrile [72]. The Menschutkin reaction is a nucleophilic
substitution reaction proceeding by an SN2 mechanism between a nitrogen base,
typically a tertiary amine, and an organic electrophile such as an alkyl halide or
sulfonate to form a quaternary ammonium salt. The transition state of this reaction
is much more polar than the reactants, thus making the reaction kinetics strongly
depend on the polarity of the surrounding medium. The Menschutkin reaction
of tributylamine (TBA) with methyl p-nitrobenzenesulfonate (MNBS) in CO2-
expanded acetonitrile at 313 K was studied (Scheme 14.19), and it is found that
at a mole fraction of CO2 > 0.60, the value of the rate constant is significantly
higher than that at a mole fraction of CO2 < 0.60. At high CO2 loading, the charge-
separated transition state of the reactant may be stabilized by acetonitrile with
decreasing polarity of the liquid phase. Another factor is that the polar solutes
would be clustered and/or aggregated as CXL increasingly becomes nonpolar.

14.3 Conclusions

Several examples described in this chapter demonstrate that CO2 can act as a
promoter/modifier in organic synthetic reactions. CO2 has different macro (physi-
cal) and micro (chemical) functions depending on the reactions and conditions
used. The reaction rate and/or the product selectivity in a liquid-phase reaction may
be improved by the actions of CO2 dissolved in the liquid phase. CO2 interacts with
organic substrates, intermediates, and/or products and changes their reactivity.
Some pressure is needed for CO2 to dissolve in a certain amount, and the pressur-
ization with CO2 is a simple procedure. When water is also present with dense-
phase CO2, the reaction mixture is more complicated, but interesting reaction
outcome would be obtained through additive and/or synergistic effects of CO2
and water. The multiphase reactions in the presence of dense-phase CO2 and/or
water are an interesting topic to be investigated in more detail, and the effects of
CO2 and water should be explained at molecular level. For practical application, the
multiphase reactions should be analyzed from the viewpoint of chemical engineer-
ing (mixing, phase behavior, mass transfer, and reaction kinetics), and the costs of
CO2 pressurization and depressurization, post-purification of water, and so on
should be considered.
14 Multiphase Catalytic Reactions in/Under Dense-Phase Carbon Dioxide. . . 385

References

1. Jessop PG, Subramaniam B (2007) Gas-expanded liquids. Chem Rev 107:2666–2694


2. Akien GR, Poliakoff M (2009) A critical look at reactions in class I and II gas-expanded
liquids using CO2 and other gases. Green Chem 11:1083–1100
3. Arai M, Fujita S, Shirai M (2009) Multiphase catalytic reactions in/under dense phase CO2.
J Supercrit Fluids 47:351–356
4. Akiyama Y, Fujita S, Senboku H, Rayner CM, Brough SA, Arai M (2008) An in situ high
pressure FTIR study on molecular interactions of ketones, esters, and amides with dense phase
carbon dioxide. J Supercrit Fluids 46:197–205
5. Yoo JS, Jhung SH, Lee KH, Park YS (2002) An advanced MC-type oxidation process – the
role of carbon dioxide. Appl Catal A Gen 223:239–251
6. Aresta M, Tommasi I, Quaranta E, Fragale C, Mascetti J, Tranquille M, Galan F, Fouassier M
(1996) Mechanism of formation of peroxocarbonates RhOOC(O)O(Cl)(P)3 and their reactivity
as oxygen transfer agents mimicking monooxygenases. The first evidence of CO2 insertion
into the O-O Bond of Rh(η2-O2) complexes. Inorg Chem 35:4254–4260
7. Nolen SA, Lu J, Brown JS, Pollet P, Eason BC, Griffith KN, Gläser R, Bush D, Lamb DR,
Liotta CL, Eckert CA (2002) Olefin epoxidations using supercritical carbon dioxide and
hydrogen peroxide without added metallic catalysts or peroxy acids. Ind Eng Chem Res
41:316–323
8. Rajagopalan B, Wei M, Musie GT, Subramaniam B, Busch DH (2003) Homogeneous catalytic
epoxidation of organic substrates in CO2-expanded solvents in the presence of water-soluble
oxidants and catalysts. Ind Eng Chem Res 42:6505–6510
9. Rathke JW, Klingler RJ, Krause TR (1991) Propylene hydroformylation in supercritical
carbon dioxide. Organometallics 10:1350–1355
10. Hemminger O, Marteel A, Mason MR, Davies JA, Tadd AR, Abraham MA (2002) Hydrofor-
mylation of 1-hexene in supercritical carbon dioxide using a heterogeneous rhodium catalyst.
3. Evaluation of solvent effects. Green Chem 4:507–512
11. Jin H, Subramaniam B (2004) Homogeneous catalytic hydroformylation of 1-octene in
CO2-expanded solvent media. Chem Eng Sci 59:4887–4893
12. Tominaga K, Sasaki Y (2004) Ruthenium-catalyzed one-pot hydroformylation of alkenes
using carbon dioxide as a reactant. J Mol Catal A Chem 220:159–165
13. Tominaga K (2006) An environmentally friendly hydroformylation using carbon dioxide as a
reactant catalyzed by immobilized Ru-complex in ionic liquids. Catal Today 115:70–72
14. Fujita S, Okamura S, Akiyama Y, Arai M (2007) Hydroformylation of cyclohexene with
carbon dioxide and hydrogen using ruthenium carbonyl catalyst: influence of pressures of
gaseous components. Int J Mol Sci 8:749–759
15. Weissermal K, Arpe HJ (1997) Industrial organic chemistry, 3rd edn. Wiley-VCH, Weinheim
16. Smith GV, Notheisz F (1995) Heterogeneous catalysis in organic chemistry. Academic,
London
17. Seki T, Grunwaldt JD, Baiker A (2008) Heterogeneous catalytic hydrogenation in supercritical
fluids: potential and limitations. Ind Eng Chem Res 47:4561–4585
18. Solinas M, Pfaltz A, Cozzi PG, Leitner W (2004) Enantioselective hydrogenation of imines in
ionic liquid/carbon dioxide media. J Am Chem Soc 126:16142–16147
19. Devetta L, Giovanzana A, Canu P, Bertucco A, Minder BJ (1999) Kinetic experiments and
modeling of a three-phase catalytic hydrogenation reaction in supercritical CO2. Catal Today
48:337–345
20. Chouchi D, Gourgouillon D, Courel M, Vital J, da Ponte MN (2001) The influence of phase
behavior on reactions at supercritical conditions: the hydrogenation of α-pinene. Ind Eng
Chem Res 40:2551–2554
21. Gallezot P, Richard D (1998) Selective hydrogenation of α, β-unsaturated aldehydes. Catal
Rev Sci Eng 40:81–126
386 H. Yoshida et al.

22. Bhanage BM, Ikushima Y, Shirai M, Arai M (1999) The selective formation of unsaturated
alcohols by hydrogenation of α, β-unsaturated aldehydes in supercritical carbon dioxide using
unpromoted Pt/Al2O3 catalyst. Catal Lett 62:175–177
23. Zhao F, Fujita S, Akihara S, Arai M (2005) Hydrogenation of benzaldehyde and
cinnamaldehyde in compressed CO2 medium with a Pt/C catalyst: a study on molecular
interactions and pressure effects. J Phys Chem A 109:4419–4424
24. Zhao F, Fujita S, Sun J, Ikushima Y, Arai M (2004) Carbon dioxide-expanded liquid substrate
phase: an effective medium for selective hydrogenation of cinnamaldehyde to cinnamyl
alcohol. Chem Commun 40:2–4
25. Wang J, Wang M, Hao J, Fujita S, Arai M, Wu Z, Zhao F (2010) Theoretical study on
interaction between CO2 and carbonyl compounds: influence of CO2 on infrared spectroscopy
and activity of CO. J Supercrit Fluids 54:9–15
26. Fujita S, Akihara S, Zhao F, Liu R, Hasegawa M, Arai M (2005) Selective hydrogenation
of cinnamaldehyde using ruthenium-phosphine complex catalysts with multiphase reaction
systems in and under pressurized carbon dioxide: significance of pressurization and interfaces
for the control of selectivity. J Catal 236:101–111
27. Liu R, Zhao F, Fujita S, Arai M (2007) Selective hydrogenation of citral with transition metal
complexes in supercritical carbon dioxide. Appl Catal A Gen 316:127–133
28. Bhanage BM, Ikushima Y, Shirai M, Arai M (1999) Multiphase catalysis using water-soluble
metal complexes in supercritical carbon dioxide. Chem Commun 35:1277–1278
29. Rappoprt Z (ed) (2007) The chemistry of anilines. Wiley, Chichester
30. Diao S, Qjan W, Luo G, Wei F, Wang Y (2005) Gaseous catalytic hydrogenation of nitroben-
zene to aniline in a two-stage fluidized bed reactor. Appl Catal A Gen 286:30–35
31. Nishimura S (2001) Handbook of heterogeneous catalytic hydrogenation for organic synthesis.
Wiley, New York
32. Li H, Zhao Q, Wan Y, Dai W, Qiao M (2006) Self-assembly of mesoporous Ni-B amorphous
alloy catalysts. J Catal 244:251–254
33. Xu R, Xie T, Zhao Y, Li Y (2007) Quasi-homogeneous catalytic hydrogenation over mono-
disperse nickel and cobalt nanoparticles. Nanotechnology 18:055602–055606
34. Höller V, Wegricht D, Yuranov I, Kiwi-Minsker L, Renken A (2000) Three-phase nitroben-
zene hydrogenation over supported glass fiber catalysts: reaction kinetic study. Chem Eng
Technol 23:251–255
35. Meng X, Cheng H, Akiyama Y, Hao Y, Qiao W, Yu Y, Zhao F, Fujita S, Arai M (2009)
Selective hydrogenation of nitrobenzene to aniline in dense phase carbon dioxide over
Ni/γ-Al2O3: significance of molecular interactions. J Catal 264:1–10
36. Meng X, Cheng H, Fujita S, Hao Y, Shang Y, Yu Y, Cai S, Zhao F, Arai M (2010) Selective
hydrogenation of chloronitrobenzene to chloroaniline in supercritical carbon dioxide over
Ni/TiO2: significance of molecular interactions. J Catal 269:131–139
37. Zhao F, Ikushima Y, Arai M (2004) Hydrogenation of nitrobenzene with supported platinum
catalysts in supercritical carbon dioxide: effects of pressure, solvent, and metal particle size.
J Catal 224:479–483
38. Zhao F, Zhang R, Chatterjee M, Ikushima Y, Arai M (2004) Hydrogenation of nitrobenzene
with supported transition metal catalysts in supercritical carbon dioxide. Adv Synth Catal
346:661–668
39. Meng X, Cheng H, Fujita S, Yu Y, Zhao F, Arai M (2011) An effective medium of H2O and
low-pressure CO2 for the selective hydrogenation of aromatic nitro compounds to anilines.
Green Chem 13:570–572
40. Lin HW, Yen CH, Tan CS (2012) Aromatic hydrogenation of benzyl alcohol and its deriva-
tives using compressed CO2/water as the solvent. Green Chem 14:682–687
41. Chanda A, Fokin VV (2009) Organic synthesis on water. Chem Rev 109:725–748
42. Yoshida H, Wang Y, Narisawa S, Fujita S, Arai M (2013) A multiphase reaction medium
including pressurized carbon dioxide and water for selective hydrogenation of benzonitrile
with a Pd/Al2O3 catalysts. Appl Catal A Gen 456:215–222
14 Multiphase Catalytic Reactions in/Under Dense-Phase Carbon Dioxide. . . 387

43. George M, Weiss RG (2002) Chemically reversible organogels via latent gelators. Aliphatic
amines with carbon dioxide and their ammonium carbamates. Langmuir 18:7124–7135
44. Xie X, Liotta CL, Eckert CA (2004) CO2-protected amine formation from nitrile and imine
hydrogenation in gas-expanded liquids. Ind Eng Chem Res 43:7907–7911
45. Fürstner A, Koch D, Langemann K, Leitner W, Six C (1997) Olefin metathesis in compressed
carbon dioxide. Angew Chem Int Ed 36:2466–2469
46. Wittmann K, Wisniewski W, Mynott R, Leitner W, Kranemann CL, Rische T, Eilbracht P,
Kluwer S, Ernsting JM, Elsevier CJ (2001) Supercritical carbon dioxide as solvent and
temporary protecting group for rhodium-catalyzed hydroaminomethylation. Chem Eur J
7:4584–4589
47. Wei HH, Yen CH, Lin HW, Tang CS (2013) Synthesis of bimetallic Pd-Ag colloids in
CO2-expanded hexane and their application in partial hydrogenation of phenylacetylene.
J Supercrit Fluids 81:1–6
48. Yoshida H, Kato K, Wang J, Meng X, Narisawa S, Fujita S, Wu Z, Zhao F, Arai M (2011)
Hydrogenation of nitrostyrene with a Pt/TiO2 catalyst in CO2-dissolved expanded polar and
nonpolar organic liquids: their macroscopic and microscopic features. J Phys Chem C
115:2257–2267
49. Reichardt C (2003) Solvents and solvent effects in organic chemistry, 3rd edn. Wiley-VCH,
Weinheim
50. Minder B, Mallat T, Pickel KH, Steiner K, Baiker A (1995) Enantioselective hydrogenation of
ethyl pyruvate in supercritical fluids. Catal Lett 34:1–9
51. Ferri D, Bürgi T, Baiker A (2002) Probing boundary sites on a Pt/Al2O3 model catalyst by CO2
hydrogenation and in situ ATR-IR spectroscopy of catalytic solid–liquid interfaces. Phys
Chem Chem Phys 4:2667–2672
52. Burgener M, Ferri D, Grunwaldt JD, Mallat T, Baiker A (2005) Supercritical carbon dioxide:
an inert solvent for catalytic hydrogenation? J Phys Chem B 109:16794–16800
53. Arunajatesana V, Subramaniama B, Hutchensonc KW, Herkes FE (2007) In situ FTIR
investigations of reverse water gas shift reaction activity at supercritical conditions. Chem
Eng Sci 62:5062–5069
54. Zeinalipour-Yazdi CD, Cooksy AL, Efstathiou AM (2007) A diffuse reflectance infrared
Fourier-transform spectra and density functional theory study of CO adsorption on
Rh/γ-Al2O3. J Phys Chem C 111:13872–13878
55. Dong LB, McVicker GB, Kiserow DJ, Roberts GW (2010) Hydrogenation of polystyrene in
CO2-expanded liquids: the effect of catalyst composition on deactivation. Appl Catal A Gen
384:45–50
56. Yoshida H, Narisawa S, Fujita S, Liu R, Arai M (2012) In situ FTIR study on the formation and
adsorption of CO on alumina-supported noble metal catalysts from H2 and CO2 in the presence
of water vapor at high pressures. Phys Chem Chem Phys 14:4724–4733
57. Arai M, Nishiyama Y, Ikushima Y (1998) Optical absorption of fine gold particles in
supercritical carbon dioxide for the characterization of solvent properties. J Supercrit Fluids
13:149–153
58. Liu R, Yoshida H, Narisawa S, Fujita S, Arai M (2012) The dispersion of TiO2 modified by the
accumulation of CO2 molecules in water: an effective medium for photocatalytic H2 produc-
tion. RSC Adv 2:8002–8006
59. West KN, Wheeler C, McCarney JP, Griffith KN, Bush D, Liotta CL, Eckert CA (2001) In situ
formation of alkylcarbonic acids with CO2. J Phys Chem A 105:3947–3948
60. Xie X, Liotta CL, Eckert CA (2004) CO2-catalyzed acetal formation in CO2-expanded
methanol and ethylene glycol. Ind Eng Chem Res 43:2605–2609
61. Chamblee TS, Weikel RR, Nolen SA, Liotta CL, Eckert CA (2004) Reversible in situ acid
formation for β-pinene hydrolysis using CO2 expanded liquid and hot water. Green Chem
6:382–386
62. Weikel RR, Hallett JP, Liotta CL, Eckert CA (2006) Self-neutralizing in situ acid catalysts
from CO2. Top Catal 37:75–80
388 H. Yoshida et al.

63. Weikel RR, Hallett JP, Liotta CL, Eckert CA (2007) Self-neutralizing in situ acid catalysis for
single-pot synthesis of iodobenzene and methyl yellow in CO2-expanded methanol. Ind Eng
Chem Res 46:5252–5257
64. Morita DK, Pesiri DR, David SA, Glaze WH, Tumas W (1998) Palladium-catalyzed carbon-
carbon bond formation in supercritical carbon dioxide. Chem Commun 1397–1398
65. Fujita S, Yuzawa K, Bhanage BM, Ikushima Y, Arai M (2002) Palladium-catalyzed Heck
coupling reactions using different fluorinated phosphine ligands in compressed carbon dioxide
and conventional organic solvents. J Mol Cat A Chem 180:35–42
66. Bhanage BM, Fujita S, Arai M (2003) Heck reactions with various types of palladium complex
catalyst: application of multiphase catalysis and supercritical carbon dioxide. J Organomet
Chem 687:211–218
67. Fujita S, Tanaka T, Akiyama Y, Asai K, Hao J, Zhao F, Arai M (2008) Impact of carbon
dioxide pressurization on liquid phase organic reactions: a case study on Heck and Diels-Alder
reactions. Adv Synth Catal 350:1615–1625
68. Weinstein RD, Renslo AR, Danheiser RL, Tester JW (1999) Silica-promoted Diels-Alder
reactions in carbon dioxide from gaseous to supercritical conditions. J Phys Chem B
103:2878–2887
69. Ikushima Y, Saito N, Arai M (1992) Supercritical carbon dioxide as reaction medium:
examination of its solvent effects in the near-critical region. J Phys Chem 96:2293–2297
70. Ford JW, Lu J, Liotta CL, Eckert CA (2008) Solvent effects on the kinetics of a Diels-Alder
reaction in gas-expanded liquids. Ind Eng Chem Res 47:632–637
71. Moral D, Osuna AMB, Córdoba A, Moretó JM, Veciana J, Ricart S, Ventosa N (2009)
Versatile chemoselectivity in Ni-catalyzed multiple bond carbonylations and cyclocarbo-
nylations in CO2-expanded liquids. Chem Commun 31:4723–4725
72. Ford JW, Janakat ME, Lu J, Liotta CL, Eckert CA (2008) Local polarity in CO2-expanded
acetonitrile: a nucleophilic substitution reaction and solvatochromic probes. J Org Chem
73:3364–3368

You might also like